You are on page 1of 376

Ecology, Economy and Security of Supply of the

Dutch Electricity Supply System:


A Scenario Based Future Analysis

Ecology, Economy and Security of Supply of the


Dutch Electricity Supply System:
A Scenario Based Future Analysis

Proefschrift

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. dr. ir. J.T. Fokkema,
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op donderdag 9 oktober 2008 om 12:30 uur
door

Johannes Gerhardus RDEL

Werktuigkundig ingenieur
geboren te Doetinchem.

Dit proefschrift is goedgekeurd door de promotor:


Prof. dr. ir. A.H.M. Verkooijen

Copromotor:
Dr. R.W. Knneke

Samenstelling promotiecommissie:
Rector Magnificus, voorzitter
Prof. dr. ir. A.H.M. Verkooijen, Technische Universiteit Delft, promotor
Dr. R.W. Knneke, Technische Universiteit Delft, copromotor
Prof. dr. ir. W. Dhaeseleer, Katholieke Universiteit Leuven
Prof. dr. J.P.M. Groenewegen, Technische Universiteit Delft
Prof. ir. J.P. van Buijtenen, Technische Universiteit Delft
Prof. ir. W.L. Kling, Technische Universiteit Eindhoven
Dr. F. van der Hooft, N.V. Nuon

The research described in this thesis was made possible through financial support from Nuon.

Cover design: Existing and new supply options (background: satellite image of the Netherlands
(NASA), left side: Waste Incineration (AEB), CCGT, Conventional coal, CHP CCGT; right side:
Electricity storage (KEMA), Offshore wind (Nuon), IGCC (Nuon), Solar energy (Nuon)).
The source is mentioned for all figures, diagrams and graphs which were not created by the author. In
so far as was possible permission for use was requested and acquired. The author recognizes the
intellectual property rights of rightful owners who could not be traced or contacted.
Printed by: Ponsen & Looijen b.v., Wageningen, The Netherlands
ISBN 978-90-6464-291-3
All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any
electronic or mechanical means (including photocopying, recording or information storage retrieval)
without permission in writing form the author.

Summary

Summary
PhD thesis Ecology, Economy and Security of Supply of the Dutch Electricity Supply System:
A Scenario Based Future Analysis by Hans Rdel
Energy, environment and society are three concepts that are inextricably intertwined. Society cannot
do without energy, for energy is essential to economic prosperity. However, this leads to increasing
environmental pressure because though energy intensity is declining in the GDP, energy consumption
per head of the population is still rising, largely due to greater use of electricity.
The electricity and heat used in the Netherlands are produced mainly by large-, medium- and smallscale technologies that convert predominantly fossil fuel into electricity and heat. A small percentage
comes from renewable sources, such as wind, sun and water, and geothermal and bio-energy,
including waste.
In the wake of an EU directive the Dutch electricity sector was transformed in 1998-2004 from a
regional/national utility to an international liberalized market with national (public sector) and
international privatized producers and providers. A liberalized electricity market is considered the best
guarantee for an efficient electricity supply as freedom of choice makes the market players compete
and invest in innovation. But, does a liberalized electricity market also offer guarantees for an optimal
balance in the objectives for environment, economy and security of supply, as defined in this study?
The supply of fossil fuels is finite and the larger part of the Dutch electricity supply goes hand in hand
with CO2 and, to a lesser extent, NOx and SOx emissions which affect the environment. There is a
general consensus that the environmental burden must be tempered as far as possible, in the form of
CO2 emission reductions by the electricity supply system. The government has set new targets for
energy savings, renewable energy and CO2 emissions for 2020. It is up to the market to achieve these
targets with or without guidelines for, amongst others, renewable energy sources.
This study looks at the current electricity supply situation. It shows how the system should be
structured to attain the environmental targets, it discusses the various growth parameters for the
different electricity production options, and explains the required incentives. But the environment is
only one of the study objectives. No less important are economic efficiency and security of supply of
the electricity system.
Thats why this study focuses on the following questions:
Which electricity supply scenarios are possible in the Netherlands, taking account of trends in
electricity/heat/fuel supply and demand, economic protectionism versus globalization, and the
importance of environmental aspects in a deregulated electricity market? How do these scenarios
affect the three objectives: energy savings/CO2 emission reductions, improved economic efficiency of
the electricity supply system, and maintaining a high level of security of supply? Which government
incentives would be needed?
To determine the effects of the selected scenarios on the three objectives of this study a supplydemand techno-economic model has been developed which can calculate the electricity supply for
any given year on an hour basis. On the basis of predefined scenarios the simulation model can be
used to address optimization issues. The study presupposes an electricity market that works optimally
and is impervious to external factors.
This dissertation describes the outcome of the study. It is split into ten chapters which are divided over
an introduction, five separate parts, conclusions, and recommendations. Below a short description is
given for each part.

Introduction: Evolution of the Electricity Supply


The introduction traces the gradual evolution of the electricity supply from a private initiative around
1880 to the current deregulated system with diverse market players, generation technologies and
electricity import/export. It describes the effects that this process has had on the environment, the
economy and security of supply and identifies various trends. Clearly, emissions of NOx and SO2 in
particular have risen sharply in recent decades and CO2 emissions are continuing to rise with the
growth in the demand for electricity. Mounting fuel prices have pushed up the generation costs despite
efficiency improvements. Since the deregulation of the market scarcely any new electricity capacity
has been built, resulting in a lower reserve factor and a growing dependence on imports.
Part 1: The Reference Situation for the Environment, Economy and Security of Supply
In Part 1the reference situation the electricity supply system in 2003 is analyzed in terms of the
environment, the economy and security of supply in order to define the frame of reference for the
scenario analyses. The first part of Chapter 2 presents a brief overview of national and EU
environmental guidelines and their aims. These guidelines have an effect on the market and influence
the operational deployment of the supply options. For the purposes of the scenario analyses the total
CO2 emissions from the domestic electricity supply are fixed at 63.5 Mtons, or 72.3 Mtons including
those from electricity imports. The second part of Chapter 2 sketches the economic situation and
briefly discusses the electricity markets and price trends which were visible in 2003. On the basis of
the price levels in the reference situation, the generation costs are analyzed for four electricity
generation options and conclusions are drawn about the performance of these generation options and
the deployment in peak (working) and off-peak (other) hours. Chapter 3 concentrates on security of
supply, beginning with a description of the methodologies to analyze security of supply. An overview is
provided of the supply and demand for electricity in 2003, followed by an analysis of the balance
between supply and demand and the implications for security of supply.
Part 2: Energy Generation, Transportation and Distribution in the Netherlands
Chapter 4 explains the current electricity supply system in the Netherlands. This is split into
categories: conventional and CCGT generation, CHP options and renewable sources and waste
incineration plants. After a brief description of thermodynamic cycles and possible ways of combining
them, the physical locations/installations, the used thermodynamic cycle and the fuel supply system
are explained for each category along with a brief outline of the production process. As a preface to
the discussion on CHP, the definition of heat/power as understood in this thesis is explained and the
advantages of CHP are offset against conventional generation and CCGT without heat. CHP
installations larger than 100 MWe are described at the same manner as conventional and CCGT
options. CHP installations smaller than 100 MWe are described in clusters. The renewable energy
sources are divided into wind, water, solar and bio-energy and are also described in clusters. The
physical waste incinerators are addressed with a brief description of the technology. This is followed
by a comparison of the key characteristics of the various generation options, including specific fuel
consumption, on/off times and maintenance. The influence of heat delivery on performance is also
assessed. An overview is presented of the environmental profiles of the current generation options
and the openings for the deployment of alternative fuel. The chapter ends with a description of the
building process for new generation, from planning to commissioning, and the influence on security of
supply. The transport and distribution which are needed to close the supply-and-demand chain are
discussed in Chapter 5 in relation to electricity, gas and CHP. Future developments in transport and
distribution, such as Distributed Generation and Demand Side Management, are addressed briefly
as well as low temperature district heating systems and developments in central cold systems.
Part 3: Latest Trends in the Development of Generation Options
In Part 3 a review of the latest trends in the development of generation techniques builds a bridge
between the description of the system in Part 2 and the techno-economic simulation model in Part 4.
Future developments in the categories listed in Chapter 4 conventional, CCGT, CHP, renewable
sources and waste incineration plants are described in detail along with the anticipated

vi

Summary

developments in production processes and key parameters, such as maintenance, unit size, electricity
yield and investment levels for 2012 and 2025. The developments in CO2 capturing and storage are
explained separately for conventional and CCGT generation. Capturing technology such as pre-, postand oxy-fuel combustion are identified and its effect on production processes, maintenance, unit size,
electricity yield and investment levels is assessed. CO2 transport and storage systems are touched on
but not explored in detail. The storage of electricity is a good option to absorb discrepancies between
supply and demand in the electricity supply system as such. The various systems which are currently
available and in development are mentioned. The chapter ends with six tables showing all the above
categories and developments in parameters such as unit size, electricity yield, maintenance costs and
investment costs for 2012 and 2025. These parameters together with the scenarios in Part 4 serve as
input for the model.
Part 4: Developed Scenarios and the Techno-Economic Simulation Model
Chapter 7 describes four future scenarios for the Dutch electricity supply system. First, the scenario
definitions are determined, the driving forces and uncertainties are identified and, on the basis of a
literature search, an inventory is compiled of the driving forces that can be used for this study. The
four scenarios (A, B, C and D) compiled for this study can be characterized with a matrix of
uncertainties. There are two scenarios, A and C, in which the environment and the finite nature of
fossil fuel are important, and another two scenarios, B and D, in which only the (lowest) costs are
important. The quintessential difference between the environmental scenarios and the lowest-cost
scenarios therefore lies in the nature of the economy: is it largely protectionist (regional) (A and B
scenarios) or geared to worldwide free trade (globalization) (C and D scenarios)? The four scenarios
serve as input for the techno-economic simulation model, but before the input parameters are
determined the mathematical-economic structure of the model is described in Chapter 8. The starting
points are defined: the algorithms for the modules electricity generation plants, other electricity
production and import/export are described and the required input and output parameters are
specified. Finally, the results are calculated for reference year 2003 and analyzed and evaluated to
validate the model.
Part 5: Scenario Analysis
In Chapter 9 the demand for electricity and heat in the reference year is analyzed and the load curves
are plotted, which will serve as input for the model. On the supply side, factors such as trends in the
electricity demand, price ratios for gas, coal and electricity, electricity price levels in neighboring
countries, incentives and must run deployment are important. The influence of these factors on the
development of the various production options for the above-mentioned modules is analyzed. The
subsequent load curves and the influencing factors are used to determine the parameters for the four
scenarios that feature in the model. The choice of parameters and the figures for each scenario are
explained. In Chapter 10, first the results of the model for the scenarios are graphically presented for
the environment, the economy and security of supply. Then the outcome for the three domains is
discussed for each scenario. A comparison of the figures in each case clearly shows the differences
between the scenarios. Scenarios A and C represent a low environmental burden and would allow the
government targets for the environment to be achieved. However, economic efficiency of the electricity
supply is low. Powerful incentive schemes will be needed every year, especially for protectionist
Scenario A. Also, in this protectionist environmentally-friendly scenario, a large volume of reserve
capacity will have to be built up to compensate for lapses in the supply of wind energy. In
environmentally-friendly Scenario C under worldwide free trade this is less urgent. In addition, the
system in the protectionist environmentally-friendly Scenario A will not be able to work at certain times
of the year when the electricity load is low because of the large percentage of wind energy. The heavy
environmental burden created by the lowest-cost scenarios will prevent the achievement of the
government targets, but economic efficiency is good and incentives would not be needed. The reserve
capacity requirement is low. The chapter ends with graphical representations and an analysis of the
effects of electricity storage on the electricity load curve, fuel consumption and CO2 emissions by the
electricity supply.

vii

Conclusions and Recommendations


In the Conclusions and Recommendations the research question is answered on the basis of the
scenario outcomes presented in Chapter 10. Clearly, the four scenarios differ widely in terms of their
ability to realize the objectives for the three domains. The main conclusion is that there is no single
scenario that scores well on all counts, but the results of the environmentally-friendly Scenario C
under worldwide free trade show that, by applying a smart combination of generation technologies,
such as gasification with CO2 capturing and storage, heat/power, on- and off-shore wind energy and
biomass (clean wood, waste wood and residual streams such as chicken manure) and by re-using
residual heat an option emerges which strikes a reasonable balance between the environment, the
economy and security of supply. This and the other conclusions form the basis of the
recommendations to the government on the adoption of an electricity supply system that strikes the
best possible balance between the three objectives of energy savings/CO2 emission reductions,
improved economic efficiency of the electricity supply system, and maintaining a high level of security
of supply. The growth parameters of the various production options can be determined and incentive
systems devised to achieve this balance. The simulation model is highly versatile and can be used to
work out other conceptual scenarios in the Netherlands and also in the EU.

viii

Samenvatting

Samenvatting
Proefschrift Milieu, economie en voorzieningzekerheid van de Nederlandse
elektriciteitsvoorziening: Een toekomst analyse op basis van scenarios door Hans Rdel.
Energie, Milieu en Samenleving, drie begrippen die onlosmakelijk met elkaar zijn verbonden. De
samenleving kan niet zonder energie, energie is welvaart. Dit leidt echter tot een toenemende
milieubelasting, want ondanks dat de energie-intensiteit in het BNP dalend is, blijft het energieverbruik
per hoofd toenemen voornamelijk door toename van het elektriciteitverbruik.
De elektriciteit en warmte die in Nederland wordt gebruikt, wordt voornamelijk geproduceerd met
grote-, middelgrote en kleinschalige energie conversie technologien die ingezet worden om
voornamelijk fossiele brandstoffen te converteren naar afgeleide energiedragers elektriciteit en
warmte. Een klein gedeelte van deze energiedragers wordt geproduceerd uit hernieuwbare
energiebronnen zoals wind, zon, water, geothermie en bio-energie waaronder afval.
De Nederlandse elektriciteitssector is als uitvloeisel van EU directieven in de periode 1998-2004
getransformeerd van een regionale/nationale nutsvoorziening tot een internationale geliberaliseerde
energiemarkt met nationale (publieke) en internationale geprivatiseerde producenten en leveranciers.
Een geliberaliseerde elektriciteitsmarkt wordt gezien als de beste garantie voor een efficinte
elektriciteitsvoorziening waarbij keuzevrijheid voor afnemers de nodige concurrentie teweeg brengt en
marktpartijen blijft stimuleren tot innovatie. De vraag die daarbij echter gesteld kan worden is of een
geliberaliseerde elektriciteitsmarkt tegelijkertijd garanties biedt voor een optimale balans tussen de
drie doelstellingen milieu, economie en voorzieningszekerheid. Het aanbod aan fossiele
brandstoffen is eindig n het grootste deel van de Nederlandse elektriciteitsvoorziening gaat gepaard
met milieubelastende emissies als CO2 en in mindere mate NOx en SOx. Er bestaat inmiddels een
groot draagvlak voor de verdere matiging van de milieubelasting en dan in de vorm van CO2 emissie
reductie van de elektriciteitsvoorziening. De overheid heeft daarvoor nieuwe doelstellingen vastgelegd
waarin energiebesparing, aandeel hernieuwbare energie en reductie van CO2 worden bepaald voor
2020. De wijze waarop dit bereikt moet worden wordt aan marktpartijen overgelaten al dan niet
aangevuld middels ondersteuningskaders voor bijvoorbeeld hernieuwbare energiebronnen.
Deze studie laat zien hoe uitgaande van de huidige elektriciteitsvoorziening het systeem ingericht zou
moeten worden om de gewenste milieu doelstellingen te kunnen bereiken, welke groeikaders voor de
diverse elektriciteitsproductieopties daarbij mogelijk zijn en welke stimuleringskaders daarvoor
noodzakelijk zijn. De gewenste milieu doelstelling is echter slechts n van de beoogde
doelstellingen. Evenzeer zijn de doelstellingen voor economische efficintie en de
voorzieningszekerheid van de elektriciteitsvoorziening van belang.
De onderzoeksvraag is daarom als volgt geformuleerd:
Welke scenarios zijn mogelijk voor de Nederlandse elektriciteitsvoorziening rekening houdend met
trends in vraag en aanbod van elektriciteit, warmte en brandstoffen, een economie ingesteld op
protectionisme versus globalisering en het belang van milieu aspecten in een geliberaliseerde
elektriciteitsmarkt? Wat zijn de effecten van deze scenarios op de drie doelstellingen :
besparing/reductie van CO2 emissies, verbeterde economische efficiency van de
elektriciteitsvoorziening en het handhaven van een grote mate van voorzieningzekerheid? Welke
overheid maatregelen zijn daarbij nodig om dit te ondersteunen.
Ter bepaling van de effecten die gekozen scenarios hebben op de drie in de onderzoeksvraag
geformuleerde doelstellingen is een vraag aanbod, technisch- economisch simulatiemodel
ontwikkeld waarmee de elektriciteitvoorziening voor een gekozen jaar met berekeningen op uur basis
kan worden doorgerekend. Aan de hand van vooraf gedefinieerde scenarios kunnen met
simulatiemodel optimalisatie vraagstukken beantwoord worden. Belangrijk uitgangspunt is hierbij de

ix

verondersteld dat er een perfect werkende elektriciteitsmarkt is zonder ondermijnende externe


invloeden.
De resultaten van dit onderzoek zijn beschreven in dit proefschrift. Het bevat in totaal 10 hoofdstukken
die zijn verdeeld over een introductie, 5 afzonderlijke delen en conclusies en aanbevelingen.
Hieronder volgt per deel een korte beschrijving van de inhoud.
Introductie: Ontstaansgeschiedenis van de elektriciteitsvoorziening
Beginnend als een privaat initiatief zo rond 1880 ontwikkelde zich stapsgewijs de huidige en
geliberaliseerde elektriciteitsvoorziening met diverse marktpartijen, diverse opwekopties en
import/export van elektriciteit. De effecten die deze ontwikkelingen hebben gehad op milieu, economie
en voorzieningszekerheid worden beschreven en trends worden benoemd. Duidelijk is dat de
emissies van met name NOx en SO2 de afgelopen decennia fors zijn afgenomen maar dat CO2
emissies blijven stijgen met de toename van de elektriciteitsbehoefte. De opwekkosten zijn als gevolg
van de alsmaar stijgende brandstofprijzen steeds hoger geworden ondanks brandstofefficiency
verbetering van de opwek. Sinds de liberalisering van de elektriciteitsmarkt is er nauwelijks nieuw
elektrisch vermogen bijgebouwd met als gevolg een dalende reservefactor en toenemende
afhankelijkheid van gemporteerde elektriciteit.
Deel 1: De referentie situatie voor milieu economie en voorzieningszekerheid
De situatie van de elektriciteitsvoorziening in 2003 is beschreven voor de aspecten milieu, economie
en voorzieningszekerheid en dient als referentie kader voor de scenario analyses die later volgen. In
hoofdstuk 2, het 1e deel wordt voor het milieu aspect eerst een korte inventarisatie gemaakt van
nationale- en EU reguleringskaders en de doelstelling daarvan. Deze kaders hebben uitwerking op de
markt en benvloeden de operationele inzet van de aanbod opties welke worden beschreven.Ten
behoeve van de scenario analyses is het totaal van CO2 emissies van de binnenlandse
elektriciteitsvoorziening vastgesteld op 63.5 Mton. Worden ook de CO2 emissies als gevolg van
import elektriciteit meegenomen dan wordt deze CO2 emissie 72.3 Mton. De economie wordt in
hoofdstuk 2, het 2e deel beschreven waarbij wordt kort wordt ingegaan op de aanwezige
elektriciteitsmarkten en prijs trends die in 2003 waarneembaar waren. Uitgaande van de prijsniveaus
in de referentie situatie in 2003 zijn analyses gemaakt van de opwekkosten voor een viertal
elektriciteitsproductieopties. Er worden uitspraken gedaan over performance van deze opwekopties en
de inzet in peak (werkdaguren) en offpeak (overige uren) van deze opties. In hoofdstuk 3 wordt
ingegaan op het aspect voorzieningszekerheid. Daarbij wordt begonnen met een beschrijving van de
methodieken om voorzieningszekerheid te analyseren. Na inventarisatie van vraag en aanbod van
elektriciteit in 2003 wordt vervolgens een analyse gemaakt hoe de balans is tussen vraag en aanbod
en wat de gevolgen zijn voor de voorzieningszekerheid.
Deel 2: Energie opwek, transport en distributie van energie in Nederland
Hoofdstuk 4 beschrijft de bestaande elektriciteitvoorziening in Nederland. De categorien die daarin
worden onderscheiden zijn conventionele en STEG opwek opties, warmte/kracht opwek opties en als
laatste de hernieuwbare energiebronnen en afvalverbrandingsinstallaties. Eerst wordt een korte
beschrijving gegeven van thermodynamische cycles en mogelijke combinaties daarvan. Daarna wordt
per categorie uiteen gezet welke fysieke locaties/installaties het betreft, welke thermodynamische
cycle wordt gehanteerd, hoe de brandstofvoorziening is georganiseerd en hoe het productie proces
verloopt. Als inleiding op de uiteenzetting bij warmte/kracht wordt de definitie van warmte/kracht voor
dit onderzoek vastgelegd en wordt aangetoond welke voordelen warmte/kracht heeft ten opzichte van
conventionele en STEG opwek opties zonder warmtelevering. De warmte/kracht installaties groter dan
100 MWe worden beschreven zoals bij de conventionele en STEG opwek opties. De warmte/kracht
installaties kleiner dan 100 MWe worden als cluster beschreven. De hernieuwbare energiebronnen
worden verdeeld in wind, water, zon en bio-energie opwek opties en als cluster kort beschreven. De
fysieke AVIs locaties/installaties worden benoemd met daarnaast een korte beschrijving van de
gehanteerde technologie. Vervolgens worden de karakteristieken zoals specifiek brandstofverbruik,
start- en stop tijden en onderhoudsaspecten van de diverse opwekopties vergeleken en wordt de

Samenvatting

invloed van warmtelevering op de performance bepaald. Een overzicht van de milieukarakteristiek van
de huidige opwek opties wordt gegeven en de mogelijkheden voor inzet alternatieve brandstofsoorten
benoemd. Het hoofdstuk wordt afgesloten met een beschrijving van het nieuwbouwproces voor
nieuwe opwek van de planningsfase tot commisionings fase en de invloed welke dit heeft op
voorzieningszeker. Voor het sluiten van de keten tussen vraag en aanbod is transport en distributie
van de energiedragers noodzakelijk. Dit wordt in hoofdstuk 5 beschreven voor transport en distributie
van elektriciteit, gas en warmte/koude. Bij alle transport en distributie wordt kort ingegaan op
toekomstige ontwikkelingen zoals Distributed Generation en Demand Side Management bij
elektriciteit alsmede laag temperatuur stadswarmte systemen en ontwikkelingen in centrale koude
systemen.
Deel 3: Nieuwe trends in de ontwikkeling van opwekopties
In dit deel wordt een brug geslagen tussen de systeembeschrijving van deel 2 en het technischeconomisch simulatiemodel dat beschreven wordt in deel 4 door het beschrijven van de nieuwe trends
in de ontwikkeling van de opwekopties. De toekomstige ontwikkelingen voor de in hoofdstuk 4
genoemde categorien conventionele en STEG opwek, warmte/kracht en hernieuwbare
energiebronnen en afvalverbrandingsinstallaties worden uitgebreid beschreven. Daarbij worden de
verwachtingen voor de ontwikkelingen in het productieproces en parameters zoals onderhoud, unit
grootte, elektrische rendementen en investeringsniveaus voor de perioden 2012 en 2025 benoemd.
Als apart onderdeel van de categorie conventionele en STEG opwek worden de ontwikkelingen in CO2
afvang en opslag beschreven. Afvang technieken als pre- post- and oxyfuel combustion worden
benoemd en de effecten voor het productieproces, onderhoud, unit grootte, elektrische rendementen
en investeringsniveaus worden bepaald. Transport en opslag systemen voor CO2 wordt aangestipt
maar er wordt niet diep op ingegaan. Voor de elektriciteitsvoorziening als zodanig is opslag van
elektriciteit een goede optie om discrepanties tussen vraag en aanbod van elektriciteit op te kunnen
vangen. Aan de diverse beschikbare en in ontwikkeling zijnde systemen wordt kort aandacht
geschonken. Het hoofdstuk besluit met een zestal overzicht tabellen voor alle hierboven genoemde
categorien en met daarin de ontwikkelingen van parameters als unit grootte, elektrische
rendementen, onderhoudskosten en investeringskosten voor de jaren 2012 en 2025. Deze
parameters dienen in combinatie met de afgeleide scenarios in deel 4 als input voor het
simulatiemodel.
Deel 4: Ontwikkelde scenario en het technisch-economisch simulatiemodel
In hoofdstuk 7 wordt de werkwijze beschreven hoe vier toekomstige scenarios worden afgeleid voor
de Nederlandse elektriciteitsvoorziening. Hiertoe worden eerst scenario definities bepaald, de
drijvende krachten en onzekerheden genventariseerd en wordt middels onderzoek van relevante
literatuur met betrekking tot scenarios voor elektriciteitsvoorziening een inventarisatie gemaakt van de
drijvende krachten die gebruikt kunnen worden voor deze studie. De vier voor dit onderzoek afgeleide
scenarios, A, B, C en D kunnen worden gekarakteriseerd aan de hand van de matrix van
onzekerheden waaruit twee scenarios, het A en C scenario, naar voren komen waarin milieu en
eindigheid van fossiele brandstoffen belangrijk zijn en twee scenario, het B en D scenario waarin
alleen (laagste) kosten belangrijk zijn. Onderscheidt tussen de milieuscenarios en de laagste
kostenscenariosonderling is vervolgens gelegen in de aard van de economie. Deze is in het ene
geval protectionistisch (regionaal) ingesteld, scenario A en B en in het ander geval, scenarios C en D
betreft het een wereldwijd omvattende vrije economie (globalisering). De vier scenarios dienen als
input voor het technisch-economisch simulatiemodel. Voordat echter de input parameters bepaald
worden wordt eerst de mathematische/economische structuur van het simulatie model beschreven in
hoofdstuk 8. Uitgangspunten worden gedefinieerd, de gebruikte mathematische algoritmen voor de
modules elektriciteitsproductiebedrijven, overige elektriciteitsproductie en import/export worden
beschreven en de benodigde invoer en uitvoer parameters worden benoemd. Tot slot worden de
model resultaten berekend voor het referentiejaar 2003 en worden de uitkomsten geanalyseerd en
gevalueerd ter validatie van het model.

xi

Deel 5: Analyse van scenarios


In hoofdstuk 9 wordt de elektriciteit en warmte vraag van de referentiesituatie geanalyseerd en worden
de load curves voor elektriciteit en warmte afgeleid die als input dienen voor het simulatiemodel. Voor
de aanbodzijde zijn factoren als ontwikkeling van de elektriciteitsvraag, de gas/kolen/elektriciteit prijs
verhoudingen, de prijsniveaus van elektriciteit in de ons omringende landen, stimuleringskaders en
must run inzet belangrijk. De invloed van deze factoren op de ontwikkelingen van de verschillende
productieopties voor de hierboven benoemde modules wordt geanalyseerd. Met de afgeleide load
curves en de van invloed zijnde factoren worden vervolgens de parameters bepaald voor de vier
scenarios die dienen als input voor het simulatiemodel. De keuze van de gebruikte parameters voor
het betreffende scenario wordt vervolgens toegelicht en de gehanteerde getallen worden onderbouwd.
In hoofdstuk 10 worden de model uitkomsten voor de doorgerekende scenarios uit deel 4 eerst
grafisch gepresenteerd voor de drie doelstellingen milieu, economie en voorzieningszekerheid en
vervolgens wordt per scenario ingegaan op de uitkomsten voor de drie doelstellingen. Een
getalsmatige vergelijking tussen de scenarios laat duidelijk de verschillen zien tussen de scenarios
onderling. Voor de scenarios A en C met een lage milieubelasting geldt dat de milieudoelstellingen
van de overheid kunnen worden bereikt. Echter de economische efficintie van de
elektriciteitsvoorziening voor deze scenarios is laag en vooral in het protectionistische scenario A is
een fors jaarlijks stimuleringskader noodzakelijk. Daarnaast moet in dit protectionistische
milieuvriendelijke scenario tevens een behoorlijk aandeel reserve vermogen worden opgesteld ter
compensatie van wegvallen van de invoeding van windenergie. In het milieuvriendelijke scenario C
met wereldwijde vrije markt is dit in mindere mate noodzakelijk. Tevens blijkt dat het systeem in het
protectionistische milieuvriendelijke scenario A op een aantal momenten in het jaar niet kan
functioneren bij lagere elektriciteitsbelasting als gevolg van het grote aandeel windvermogen. De
kosten-scenarios hebben een hoge milieubelasting waarmee overheidsdoelstellingen niet worden
gehaald. De economische efficintie is echter goed en stimuleringskaders zijn hier niet van
toepassing. Het aandeel noodzakelijk reservevermogen is laag. Het hoofdstuk besluit met de
grafische weergaven en analyse van de effecten van elektriciteitsopslag op de elektriciteit load curve
en de gevolgen van elektriciteit opslag voor brandstofverbruik en CO2 emissie van de
elektriciteitsvoorziening.
Conclusies en aanbevelingen
In de conclusies en aanbevelingen wordt de onderzoeksvraag beantwoord aan de hand van de
scenario uitkomsten uit hoofdstuk 10. Duidelijk blijkt dat er voor de vier gedefinieerde scenarios grote
verschillen optreden voor de drie doelstellingen. Belangrijkste conclusie is dat er niet n scenario is
dat voldoet aan alle drie doelstellingen. De uitkomsten van het milieuvriendelijke scenario C met
wereldwijde vrije markt laat echter zien dat er middels slimme combinatie van verschillende
opwektechnologien zoals vergassingstechnologie met CO2 afvang en opslag, warmte/kracht, on- en
off shore wind energie en de inzet van biomassa (schoon hout, afval hout en reststromen als
kippenmest) en hergebruik van restwarmte een variant ontstaat die in redelijke mate voldoet aan de
optimale balans tussen milieu, economie en leverzekerheid. Deze conclusie en de overige conclusies
leiden tot aanbevelingen voor de keuzes die de overheid kan maken om te komen tot een
elektriciteitsvoorziening die voldoet aan de meest optimale balans voor de drie geformuleerde
doelstellingen besparing/reductie van CO2 emissies, verbeterde economische efficiency van de
elektriciteitsvoorziening en het handhaven van een grote mate van voorzieningzekerheid. Tevens
kunnen de groei kaders van de diverse productieopties worden vastgelegd en stimuleringskaders
worden bepaald waarmee deze optimale balans bereikt kan worden. Tenslotte blijkt het
simulatiemodel zeer geschikt te zijn om ook andere mogelijke scenarios door te rekenen voor de
Nederlandse elektriciteitsvoorzienig maar ook op EU schaal.

xii

Contents

Contents
Summary ..................................................................................................................................................v
Samenvatting........................................................................................................................................... ix
Contents ............................................................................................................................................... xiii
Chapter 1 Introduction .........................................................................................................................1
1.1
Evolution of the Dutch Electricity Sector........................................................................................3
1.2
Effects of Evolution on Environment, Economics, and Security of Supply....................................8
1.2.1 Environment .........................................................................................................................8
1.2.2 Economics..........................................................................................................................11
1.2.3 Security of Supply ..............................................................................................................12
1.2.4 Observations ......................................................................................................................13
1.3
Research Objective and Questions .............................................................................................15
1.3.1 Research Objective............................................................................................................15
1.3.2 Problem Definition..............................................................................................................16
1.3.3 Research Questions ..........................................................................................................17
1.3.4 Scope of the Research ......................................................................................................18
1.4
Thesis Overview ..........................................................................................................................19
References ..................................................................................................................................20

Part 1

Description of the Reference Situation in 2003

Chapter 2 Environment and Economics ..........................................................................................25


2.1
Introduction .................................................................................................................................25
2.2
Environmental (CO2) Performance .............................................................................................25
2.2.1 CO2 Emissions Electricity Production Sector ....................................................................26
2.2.2 Effects of Choices in Electricity Production Options on CO2 Emissions ...........................27
2.3
Economic Performance ..............................................................................................................27
2.3.1 Electricity Market Pricing in 2003 ......................................................................................28
2.3.2 Anticipated Trends in Peak and Off-Peak Prices for 2003 ...............................................30
2.4
Summary ....................................................................................................................................32
Chapter 3 Security of Supply ............................................................................................................33
3.1
Current Method of Operation ......................................................................................................33
3.2
Methods Applied to Analyze Security of Supply .........................................................................33
3.3
Demand for Electricity .................................................................................................................35
3.4
Electricity Supply ........................................................................................................................37
3.5
Matching Supply and Demand ...................................................................................................38
3.6
Summary ....................................................................................................................................40
References ..................................................................................................................................41

Part 2

Description of the System

Chapter 4 Energy Supply Chain and Supply Options .....................................................................45


4.1
Introduction ................................................................................................................................45
4.2
Energy Supply Chain Overview ...............................................................................................45
4.3
Definition Supply Side ...............................................................................................................45
4.3.1 Main Categories ...............................................................................................................45
4.3.2 Basic and Combined Cycles in Energy Conversion Technology ......................................47
4.3.3 Definition of Input and Output Parameters for the Model .................................................48
4.4
Conventional and Combined Cycle Units ...................................................................................49
4.4.1 Conventional Coal-Fired Units ..........................................................................................49

xiii

4.5

4.6

4.7
4.8
4.9
4.10

4.4.2 Conventional Gas-Fired Units ...........................................................................................52


4.4.3 Hot Windox Combined Cycle Gas Turbine (HW CCGT) .................................................53
4.4.4 Nuclear ..............................................................................................................................54
4.4.5 Peak-Load Units ...............................................................................................................57
4.4.6 Integrated Gasification Combined Cycle (IGCC) .............................................................58
4.4.7 Combined Cycle Gas Turbine (CCGT) ............................................................................60
Combined Heat and Power ........................................................................................................64
4.5.1 Specific Characteristics of CHP ........................................................................................64
4.5.2 Heat Supply Pattern; CHP Back-Up / Peak-Demand Facilities ............................................65
4.5.3 Combined Heat and Power Units for Industrial Heating ...................................................67
4.5.4 Combined Cycle Gas Turbine District Heating (CCGT and GT+HRSG DH) ...................76
4.5.5 Small-Scale Combined Heat and Power (Gas Engines) .................................................79
Sustainable Options ....................................................................................................................82
4.6.1 Wind Energy ......................................................................................................................82
4.6.2 Hydro Power .....................................................................................................................83
4.6.3 Photo-Voltaic Solar Energy................................................................................................84
4.6.4 Bio-Energy ........................................................................................................................84
Technical Conditions for Unit Deployment .................................................................................87
Environmental Situation ..............................................................................................................89
Security of Supply .......................................................................................................................91
Summary .....................................................................................................................................92

Chapter 5 Transport and Distribution of Energy .............................................................................93


5.1
Introduction .................................................................................................................................93
5.2
Electricity Transmission and Distribution Grids ..........................................................................93
5.2.1 Introduction .......................................................................................................................93
5.2.2 Transport and Distribution of Electricity .............................................................................94
5.2.3 Structure of the Transmission Grids ..................................................................................95
5.2.4 Structure of the Distribution Grids .....................................................................................95
5.2.5 Cross-Border Connections ................................................................................................96
5.2.6 Near Future Developments ............................................................................................ 100
5.3
Gas Transport and Distribution Grids ...................................................................................... 103
5.3.1 Introduction .................................................................................................................... 103
5.3.2 Transport and Distribution of Natural Gas ..................................................................... 103
5.3.3 Storing Natural Gas ....................................................................................................... 106
5.3.4 Near Future Developments ............................................................................................ 108
5.4
Heat Transport and Distribution Grids ..................................................................................... 109
5.4.1 Introduction ..................................................................................................................... 109
5.4.2 Transport and Distribution of District Heat...................................................................... 109
5.4.3 Near Future Developments............................................................................................. 110
5.4.4 District Heating and District Cooling ............................................................................... 112
5.5
Summary .................................................................................................................................. 113
References ............................................................................................................................... 114

PART 3 Supply Trends


Chapter 6 Technological Trends in Supply Options .................................................................... 121
6.1
Introduction .............................................................................................................................. 121
6.2
Conventional and Combined Cycle Technology ..................................................................... 121
6.2.1 Clean Coal Technologies (CCTs) ................................................................................. 121
6.2.2 Nuclear Technologies .................................................................................................... 126
6.2.3 Integrated Gasification Combined Cycle (IGCC) .......................................................... 136
6.2.4 Combined Cycle Gas Turbine (CCGT) ......................................................................... 138
6.2.5 Carbon Capture & Storage ............................................................................................ 145
6.3
Combined Heat and Power ..................................................................................................... 153
6.3.1 Combined Heat and Power Units for Industrial and District Heating ............................. 153
6.3.2 Small-Scale Combined Heat and Power (Gas Engines) .............................................. 155
6.3.3 Small-Scale Combined Heat and Power ( WKK) ........................................................ 156
6.4
Sustainable Options ................................................................................................................. 159

xiv

Contents

6.5

6.6

6.4.1 Wind Energy ................................................................................................................... 159


6.4.2 Hydro Power .................................................................................................................. 160
6.4.3 Solar Energy .................................................................................................................. 160
6.4.4 Bio-Energy ..................................................................................................................... 161
6.4.5 Electricity Storage .......................................................................................................... 162
Overview Technologies for Future Scenarios ......................................................................... 163
6.5.1 Conventional Units ......................................................................................................... 163
6.5.2 Combined Heat and Power ............................................................................................ 163
6.5.3 Sustainable Options ....................................................................................................... 164
Summary .................................................................................................................................. 165
References ............................................................................................................................... 166

Part 4

Dutch Electricity Demand and Supply Scenarios and Simulation


Model

Chapter 7 Scenarios for the Dutch Electricity Supply.................................................................. 175


7.1
Introduction ............................................................................................................................... 175
7.2
Scenarios.................................................................................................................................. 175
7.2.1 Definitions ....................................................................................................................... 175
7.2.2 Scenario Development.................................................................................................... 176
7.2.3 Scenarios for the Dutch Energy and Fuel Supply in the Literature................................. 177
7.3
Developing Scenarios for the Model......................................................................................... 186
7.3.1 Introduction ..................................................................................................................... 186
7.3.2 Driving Forces and Uncertainties.................................................................................... 186
7.3.3 Horizons .......................................................................................................................... 188
7.3.4 Narrative/Storyline Scenarios ......................................................................................... 189
7.4
Summary .................................................................................................................................. 192
Chapter 8 Dutch Electricity Supply Model ..................................................................................... 193
8.1
Introduction ............................................................................................................................... 193
8.2
Objectives, Delineation, and Simplification of the Model.......................................................... 193
8.3
Structure of the Model .............................................................................................................. 195
8.3.1 Introduction ..................................................................................................................... 195
8.3.2 Short Review of Energy Models, Comparison with Our Model....................................... 195
8.3.3 Model Structure and Sub-Model Description .................................................................. 197
8.3.4 Model Input and Output .................................................................................................. 198
8.3.5 Demand and Supply Balance ......................................................................................... 199
8.4
Mathematical/Economic Structure of the Model....................................................................... 203
8.4.1 Economics of Electricity Generation Companies, Unit Commitment/Dispatch............... 203
8.4.2 Economics of Other Electricity Producers, Combined Heat/Power, Waste Incinerators,
Non-Fossil ................................................................................................................................ 211
8.4.3 Economics of Import and Export..................................................................................... 224
8.5
Validating the Model ................................................................................................................. 225
8.6
Summary .................................................................................................................................. 226
References ............................................................................................................................... 227

Part 5

Input Model and Analysis of Scenario Outcomes

Chapter 9 Input Model for the Developed Scenarios.................................................................... 233


9.1
Introduction ............................................................................................................................... 233
9.2
Model Input: Demand Side ....................................................................................................... 233
9.2.1 Introduction ..................................................................................................................... 233
9.2.2 Electricity Demand .......................................................................................................... 233
9.2.3 Electricity Load................................................................................................................ 234
9.2.4 Heat Demand .................................................................................................................. 236
9.2.5 Heat Load........................................................................................................................ 237
9.3
Model Input: Supply Side.......................................................................................................... 238

xv

9.4

9.5

9.3.1 Introduction ..................................................................................................................... 238


9.3.2 Factors that Influence Supply Trends ............................................................................. 239
9.3.3 Electricity Production Companies ................................................................................... 242
9.3.4 Other Electricity Producers ............................................................................................. 247
9.3.5 Import/Export................................................................................................................... 259
Filling in the Model for the Developed Scenarios ..................................................................... 264
9.4.1 Introduction ..................................................................................................................... 264
9.4.2 Demand and Supply Side ............................................................................................... 264
9.4.3 Operational, Dispatch and Fixed Cost ............................................................................ 267
9.4.4 Financial Parameters ...................................................................................................... 271
9.4.5 Stimulation Measures ..................................................................................................... 273
9.4.6 Parameters Not Taken into Account............................................................................... 273
Summary .................................................................................................................................. 275

Chapter 10 Analysis of Scenario Outcomes.................................................................................. 277


10.1 Introduction ............................................................................................................................... 277
10.2 Model Results for the Developed Scenarios ............................................................................ 277
10.2.1 Environment .................................................................................................................. 277
10.2.2 Economics..................................................................................................................... 278
10.2.3 Security of Supply ......................................................................................................... 280
10.3 Analysis of the Results for 2025 ............................................................................................... 280
10.3.1 Scenario Results ........................................................................................................... 280
10.3.2 Scenario Results Compared ......................................................................................... 284
10.4 Effects of Electricity Storage for Scenario C ............................................................................ 286
10.5 Summary .................................................................................................................................. 288
References ............................................................................................................................... 289
Chapter 11 Overall Conclusions ..................................................................................................... 295
11.1 Summary .................................................................................................................................. 295
11.2 Summarizing Conclusions ........................................................................................................ 298
11.3 Recommendations.................................................................................................................... 298

Appendices
Appendix A .......................................................................................................................................... 303
Appendix B .......................................................................................................................................... 305
Appendix C .......................................................................................................................................... 307
Appendix D .......................................................................................................................................... 323
Appendix E .......................................................................................................................................... 325
Appendix F .......................................................................................................................................... 327
Nomenclature ...................................................................................................................................... 351
Dankwoord .......................................................................................................................................... 355
Curriculum Vitae .................................................................................................................................. 357

xvi

Introduction

xviii

Chapter 1 Introduction

Chapter 1

Introduction
In the past hundred years, energy supply in the Netherlands and neighboring countries has played a vital
role in ensuring that the wheels of society have continued to run smoothly and has made a significant
contribution to our standard of living. The use of energy in general and of electricity and heat in particular is
therefore bound up with the development of our society and welfare (see Figure 1.1).
180

Prosperity
160

Electricity use

140

Index (1992 = 100)

120

Energy use

100

80

60

40

J/

GDP/cap

100 = 14,759 /cap

GWh

100 = 85,874 GWh

PJ

100 = 2,802 PJ

PJ/GDP

100 = 0.0125 10^9 J/

20

0
1992

1994

1996

1998

2000

2002

2004

Figure 1.1: Development prosperity and energy use in the Netherlands (Source: CBS Statline)
After World War II the energy use per head of the population has increased ever since. Energy use was
booming at the end of the 60s and early 70s, a steep decrease occurred at the end of the 70s and early
80s and after that period, energy use still increases but with a smaller slope (see Figure 1.2).
250
First oil crisis (1973)

Second oil crisis (1979)

200

GJ/cap

Exponential prosperity growth

150

100

50

0
1945

1950

1955

1960

1965

1970

1975

1980

1985

1990

1995

2000

2005

Figure 1.2: Development in total energy requirement per capita in the Netherlands (Source: CBS Statline)

The energy intensity in gross domestic product is declining (see Figure 1.1) but with the influences of
energy on the development of our society and welfare and the continuously growing demand for energy,
particularly electricity, makes the availability of a reliable energy supply system even more important.
The energy supply system in the Netherlands comprises:
Supply systems for the energy carriers natural gas (with main- and regional transport systems),
coal, oil, uranium, waste, and biomass.
Energy conversion technologies, mainly large-, medium-, and small-scale which are used to
convert the energy carriers above into derivative energy carriers electricity, process-, and district
heat (only a small part of the energy carrier electricity is generated by converting renewable energy
sources such as wind, solar, and hydro).
The energy carriers electricity, process heat, and district heat are supplied to large industrial users,
small and medium-sized businesses, and consumers by means of electricity transport and
distribution grids, process heat distribution grids, and district heat transport and distribution grids.
Although energy conversion technologies are being improved continuously the conversion of fossil fuels
still involves substantial energy loss and environmentally harmful emissions. The deployment of the energy
supply system to provide the required energy services, the sustainability level of energy generation, and
continued security of supply come at a price which has to be borne by society as a whole. Therefore,
creating the right balance between these three pillars environment, economics and security of supply is
of eminent importance not only to reduce the environmental burden as much as possible but also to
maintain an affordable and reliable energy supply system.
The European Union was quick to recognize the crucial importance of energy in the operations of the EU
as a whole and for the competitiveness of the European economies. As a consequence, the supply of
energy carriers such as electricity and gas now is part of the internal market - an area without internal
frontiers in which the free movement of goods, persons, services and capital is ensured - in accordance
with Article 14 of the EC Treaty. The members of the European Union believe that an efficient internal
European market is important to the attainment of EU objectives, to leverage expansion opportunities by
admitting new member states, and to strengthen the economic position of the EU with an eye to the ageing
population.
Directives 96/92/EC and 98/30/EC address the realization of an internal electricity and gas market
respectively. The ultimate aim is to create an integrated internal competitive energy market in the European
Union and safeguard continuity in the energy supply. On December 19, 1996 the European Parliament and
the Council of the European Union issued Directive 96/92/EC setting out common rules for the internal
electricity market. The directive consists of a full set of measures aimed at realizing an entirely free
electricity market for the benefit of European consumers and endeavors to lay the foundations for fair and
honest competition. This directive was later replaced by Directive 2003/54/EC, which the European
Parliament and the Council issued on June 26, 2003. In response to a request from the European Summit
in Lisbon, the second directive sharpened the rules in order to be sure of realizing the objective of the
original directive 96/92/EC. The member states are now obliged to take the necessary steps to ensure that
certain objectives are clearly realized by January 1, 2007, such as protection for vulnerable buyers, basic
consumer rights, and economic and social cohesion. The EU has also issued or prepared some
supplementary guidelines for certain sub-topics relating to electricity.
The Dutch government has decided to implement this EU directive by liberalization of the electricity and
gas sector. Prime importance is accorded to freedom of choice for the buyer, as this is expected to
stimulate competition between the providers and thereby enhance cost-effectiveness. Other anticipated
benefits are lower electricity prices (before taxes), a wider choice of product packages, and improved
service. Liberalization could also bolster the competitive position on the international market (Ministry of
Economic Affairs, 2000). The Electricity Act of 1998 sets the parameters for market forces in the Dutch
electricity sector. Directive 2003/54 and a number of implementation problems have already led to four
amendments to the act and more are being drafted.

Chapter 1 Introduction

The EU decision heralded a major change for the electricity sector 1, which had always been organized
regionally and nationally as a monopoly and had evolved in tandem with changing circumstances.
The effects were compounded by the fact that there was scarcely any insight into how far the electricity
service could operate as a technological system in a liberalized market, the organization of the market
itself, the possible effects on the environment, economics and security of supply, and the potential behavior
of the market players (Boisseleau, 2004). Now that tentative experience has been gained and some
politicians and market players are raising questions about market operations, there is a distinct risk that the
age-old doctrinaire debate between the supporters and opponents of a liberalized market will be rekindled.
To add to the complications, there are no clear-cut criteria for gauging the market performance. This
situation can only be resolved by returning to the questions that should have been raised at the start of the
debate: What is technologically feasible with the subsequent implications for the balance between
environment, economics, and security of supply? What are the governments aims and how can they best
be realized in the liberalized electricity market?
The following points are crucial to a clear understanding of the technological potential and limitations
of the electricity system as a whole:
1. As a product, electricity is a 100% commodity, offering exactly the same quality to all customers. All
producers and buyers of electricity are physically interconnected via electricity transport and
distribution grids. This means that, in physical terms, electricity is never supplied directly by one
specific producer to one specific buyer and that the actions of one producer or buyer or the grid
operator (investment decisions, business operations, availability, trends in demand) will exert an
influence across the whole system.
2. As it is almost impossible at present to store electricity in large quantities, the demand must be
met as it arises. The electricity demand fluctuates strongly in the course of 24 hours and also
between working days and weekends.
3. Energy conversion technologies are incomparable in terms of applied technology, availability, cost
price of derivative energy carriers, operations, storage potential, and effects on the environment,
economics, and security of supply.
The evolution of the electricity sector in the Netherlands from its inception to the start of liberalization in
1998 is traced below in order to create a starting point for the research questions.

1.1

Evolution of the Dutch Electricity Sector

The early years


Electricity was first introduced in the Netherlands around 1880 as a private initiative. In its first main
market, lighting, it became the arch rival of stadsgas (coal gas). Though it won the battle for lighting,
electricity and stadsgas have competed against each other for various usages ever since. The
stadsgas set-up served as the role model for electricity in the early days, especially in terms of
functioning as a public utility. Electricity production was soon taken over by Public Utilities, which were
owned by (local) authorities, which held a monopoly in the areas allocated to them. The first municipal
energy board was founded in Rotterdam in 1895. These utilities were originally municipal or regional
organizations. The State first intervened in 1904 by setting up a Government Commission for Electric
Cables. The remit of this commission was to ascertain, with due respect for public safety, the legal
conditions under which electric lines and cables would be laid and used, including the accompanying
legal relationships. The commission could not deliver a satisfactory result and merely concluded that a
concession system was needed. Responsibility still rested, however, with the local authorities.
Ever since electricity was first introduced, some users have catered for their own needs by deploying
their own installations. Most of them have such high electricity and heat requirements whereby
residual heat can be used in industrial processes that it is more cost-effective for them to install their
own cogeneration system. As such installations were not exactly encouraged by the electricity boards,
swaps of surpluses and shortages with the public grid were usually kept to a minimum.
1

In the following of this research, we shall concentrate on the electricity sector.

Pretty soon, the benefits of economies of scale were discovered thanks, on the one hand, to
technological progress in electricity generation methods and grid infrastructure and, on the other hand,
to the fact that linked systems require a lower reserve capacity and are more reliable than separate
systems. The need for economies of scale further increased due to the concessions granted via the
provinces. The government came under mounting pressure to intervene and passed the Electricity Act
in 1938. The main thrust of this act was to secure optimal control over electricity. However, only the
technical aspects of the act came into effect. Further responsibility for the electricity supply was left to
the provinces and the municipalities.
The post-war period until 1970, follow the economic growth
After World War II the electricity companies realized that the only way to avoid more government
intervention was to work together more closely. Until then cooperation had been limited to informal
consultations and alignment through the Vereniging van Directeuren van Elektriciteitsbedrijven in
Nederland (VDEN / association of directors of electricity companies in the Netherlands). This led to the
establishment of N.V. Samenwerkende Electriciteits-Productiebedrijven (SEP / a partnership between
electricity producers) in 1948 and the Vereniging van Exploitanten van Elektriciteitsbedrijven in
Nederland (VEEN / association of electricity operators in the Netherlands) in 1952. SEP played a
coordinating role in the production and was responsible for import/export and production planning (the
Electricity Plan). VEEN coordinated the distribution. A system of self-regulation evolved where central
and local government had only marginal influence.
The advent of natural gas (after the discovery in Slochteren on July 22, 1959), the introduction of
petroleum and nuclear energy, and the mine closures made energy a key focus of government
attention (Vlijm, 2002), a prime example being the centralization of the gas supply via the Gasunie
(founded on April 6, 1963). Strong opposition from the electricity sector put paid, for the time being, to
any further attempts by the government to increase its influence in this sector. There was, however, a
spectacular rise in the use of natural gas to generate electricity, the argument being that the monetary
value of gas would be modest if, as expected, nuclear energy were deployed on a large scale.
The governments role went no further than setting some tariff conditions via the Prijzenbeschikking
Elektrische Energie 1952 (pricing agreement for electrical energy) and negotiations on the choice of
fuel.

Figure 1.3: Conventional coal-fired power plant Centrale Gelderland at Nijmegen in 1959 (Source:
Techniek in Nederland in de twintigste eeuw, 2000)

Chapter 1 Introduction

The 1970s, advance of gas, increasing security of supply


In the 1970s further economies of scale led to the continued concentration of electricity production in
nine provincial electricity boards (Groningen/North Drente, Friesland, Overijssel/South Drente,
Gelderland, Utrecht, North Holland, Zeeland, North Brabant and Limburg) and five municipal/regional
boards (Amsterdam, Rotterdam, The Hague, Leiden and Dordrecht). These fourteen electricity boards
were also the distributors and the joint owners of SEP. With the general partnership agreement of
1970 SEP expanded its originally limited role as coordinator of the interconnections and exchanges by
becoming owner and manager of the national 380kV interconnected grid built in the 1960s and 1970s,
including the connections with other countries, and played a greater role in the national E Plan for
mapping out the required production capacity and the internal offsetting of surpluses against
shortages. The E Plan was used to determine the units that would be decommissioned on the basis of
the anticipated demand, the new capacity that would be built and by whom, and the investments in the
interconnected grid. As of 1975 this E Plan had to be approved by the government on the basis of a
covenant between the government and SEP.
Meanwhile, a great many distribution companies (mostly municipal or regional) continued to exist (see
Figure 1.5). Government influence was still limited because most of the responsibility still rested with
the provinces and municipalities. It was the first oil crisis2 which was largely responsible for making the
government realize the need for an overall energy policy. This resulted in the First Energy Policy
Document in 1974, in which the key words were energy-saving and diversification. The aim was to
limit the deployment of natural gas and increase the deployment of coal and nuclear energy. To ease
the transition a building program was started for the storage of oil. That way, gas-fired power plants
that used oil as secondary fuel would be able to switch to oil entirely. A number of gas-fired power
plants were also converted for coal. At that time, almost all power plants were equipped for two types
of fuel (so-called dual firing). Oil was the second fuel for a few coal-fired power plants and for most
gas-fired power plants, while gas was the second fuel for some coal-fired power plants.

Figure 1.4: Conventional gas-fired power plant Flevocentrale (Source: PGEM, 1988)
2

Together with the United States, Denmark and some other countries, the Netherlands was directly targeted
through an oil embargo in 1973, because of their outspoken friendly relations with Israel. The embargo was
eventually terminated in 1974.

The 1980, consolidation energy companies


Before all planned storage tanks for oil were ready, a decision was taken in 1982 to again maximize
the deployment of natural gas in the hope of increasing the national revenue from natural gas. The
1980s were characterized by:
1. Further strengthening of the position of SEP by the introduction of the LEO optimization
algorithm (Landelijke Economische Optimalisatie / national economic optimization on the basis
of marginal costs) for determining the deployment of electricity production units, the cost
pooling, and responsibility for the Electricity Plan. SEP aspired towards a position that was as
independent as possible of the government.
2. The distributors ambition to retain influence within SEP: this led to further concentration of
electricity production in the mid-1980s. It began with the formation of EPON, which united the
electricity production of the provincial electricity boards of Gelderland and Overijssel. Soon, it
was joined by the electricity production of the provinces of Friesland and Groningen. This was
followed by even further concentration and a total separation of electricity production and
distribution. Four electricity production companies emerged (EPON, UNA, EZH and EPZ), a
few large distributors (Nuon, Essent, Eneco, Delta) and a small contingent of minor distributors
(see Figure 1.5).
250
Large-scale production
Electricity + Gas + Heat
Electricity + Heat

200

Electricity + Gas
Heat
Gas
150

Electricity

100

50

0
1970

1975

1980

1985

1990

1995

2000

2005

Figure 1.5: Energy companies in the Netherlands 1970 - 2005 (Source : EnergieNed)
3. The governments aim to increase its influence over the sector by centralizing municipal and
regional powers: the CoCoNut Report (Commissie Concentratie Nutsbedrijven / commission
for consolidation of public utilities) of 1980 proved crucially important here as it made clear
recommendations on the role of SEP and the influence of the government. The governments
intention was to define a legal framework for its relationship with the electricity sector to
replace the partially non-operational act of 1938 and the covenant with SEP of 1975. The aim
was for the sector to operate as much as possible as a commercial enterprise, with the
government at arms length. Opposition from the sector led to alterations to the original plans.
4. The Electricity Act of 1989 went into effect. Its main provisions were as follows:
a. SEP became the negotiating partner of the Ministry of Economic Affairs for the
planned new electricity production capacity, environmental policy, the national basic

Chapter 1 Introduction

rate (the LBT, landelijk basistarief) for the supply of electricity from SEP to the
distributors, and the fuel diversification. The structure of the SEP partnership
agreement would remain intact and the distributors would get their supplies from the
producers. SEP had to reach agreement with VEEN on the E Plan and the rates
before they were submitted to the government for approval.
b. The Ministry of Economic Affairs, in dialog with VEEN, set maximum rates for end
users.
c. Electricity production and distribution were separated by law, though the distributors
were allowed to build units of < 25 MW. A process of far-reaching horizontal
integration and concentration emerged within distribution.
d. Import and export remained a SEP monopoly with exceptions for specific bulk
consumers who were allowed to import electricity themselves.
5. The advent of cogeneration: as early as 1978 the Algemene Energie Raad (AER / General
Energy Council) had identified key opportunities for saving energy by means of cogeneration
and heat distribution. At first the energy companies were not particularly interested in
cogeneration, preferring to focus on efficiency improvements at power plants. This changed
gradually, especially when the government initiated a policy to stimulate cogeneration.
The 1990s on the eve of a liberalized market
Electricity as a public utility led to the following role allocation for the actors in the 1990s:
Production and Transport
The large-scale electricity production was concentrated mainly in the four electricity production
companies, which actually worked together as one via the SEP subsidiary. This gave SEP
considerable influence. Besides being the owner and manager of the nationwide 380kV transport grid
and, in the meantime the 220kV grid as well, it was responsible for exchanges with other countries,
the deployment of production units based on the LEO optimization algorithm to minimize costs,
pooling the costs of production and the associated national basic rate (NBR) for delivery by producers
to distributors, the Electricity Plan with the allocation and supervision of new building units, the
conversion of units, environmental measures and the decommissioning of units, and the purchase of
fuel.
The electricity production companies owned the power plants and were also responsible for funding
them and making them available. In return, they received standard payment. They were allowed to
add the difference between this standard payment and the real costs to the national basic rate. This
supplement plus the national basic rate resulted in the regional basic rate (RBR), which was applied to
deliveries that went from the pool via the production companies to the distributors in the region. The
supplement to the national basic rate was only a fraction of the national basic rate itself, so the
purchase price was more or less equal for all the distributors. In fact, it was not, strictly speaking, a
rate and was more like a clearance system. The fixed part of the national basic rate was determined
annually and cleared on the basis of its share in the maximum national peak demand, while the fuel
component was determined once a quarter depending on developments in the fuel prices. The costs
of the interconnected grid were allocated according to a separate system based on the share of the
buyers in the maximum national demand peak plus several additional provisions.
Distributors
The distributors were responsible for the transport (150 kV and lower) and distribution of electricity, the
distribution of natural gas, the transport and distribution of heat, and for supplying the buyers, apart
from a small group of energy-intensive businesses. These buyers were granted subsidized rates
through the special business policy of SEP and the so-called Potjesgas (leftover gas) of the Gasunie
in order to secure their competitive position on the international market. As mentioned, the distributors
had no influence on the purchase price. Pricing policy was free within the limits set by the Ministry of
Economic Affairs. The distributors were also expected, via the MAP, to implement the environmental
policy for electricity and heat in relation to the buyers. The required funding became available through
a permitted MAP levy on end-user gas and electricity rates. Levels of independence varied widely
from company to company according to the legal relationship with the owners, the local and regional
authorities.

Buyers
The prices were public and were the same for equal buyers from the same distributor. There were
differences between the rates of the different distributors, but these were generally felt to be
reasonable, except by a few bulk consumers. This was partly due to the fact that, after the mid-1980s,
natural gas became fairly cheap, making the Dutch rates relatively low in an international perspective
and making the energy bill just a limited part of the overhead costs for most buyers. In addition, natural
gas was relatively clean and the energy companies worked hard on the realization of cogeneration
projects, heat distribution projects, energy saving projects, and on cleaning flue gas in association with
and at the request of the authorities. In addition, a number of large energy-intensive industries
(exceptional bulk consumers) were granted subsidized electricity or gas rates to help them compete in
the international arena. This arrangement was introduced because self-importation rarely worked in
practice as there was no effective international arrangement for Third Party Access.
There were regulated rates for self-generators that returned electricity to the public grid.
Government
The role of the national government consisted mainly of approving the electricity plan and setting the
maximum rates for end users. By means of legislation and regulations, the national government,
acting in association with the local and regional authorities and the energy companies, exerted a
strong influence on the planological, air, water, economic and environmental requirements. These
instruments directly empowered the government to influence the balance between environment,
economics, and security of supply. The public utilities operated more or less independently within
these parameters.
The Third Energy Policy Document (Ministry of Economic Affairs, 1995) concretized the evaluation of
the 1989 Electricity Act. Its central themes were market forces and sustainability. This led, in
combination with the EU directive on the deregulation of the electricity market, to the Electricity Act of
1998, in which the government took a step back and the emphasis shifted to market forces.
The 2000s, liberalization of the electricity market
The liberalization of the electricity market by the Electricity Act of 1998 marked the end of energy as a
public service and sparked a profound change throughout the electricity sector. Within just a few years
the monopolistic centrally-run system had been transformed into fully liberalized production and supply
markets. Regulation for the grids of the distributors is handled by the Dutch energy regulator, the Dte.
Meanwhile, a great many changes were taking place in the ownership of energy companies. Three of
the four electricity production companies were taken over in 1999 by large foreign companies: EZH
had been acquired by E.on (Germany), EPON by Electrabel (Belgium), and UNA by Reliant (USA).
Later, in 2003, Reliant sold its power plants back to Nuon, most of the EPZ power plants were
transferred to the principal shareholder Essent. In 2007 the government decided that integrated
distribution companies have to be unbundled (latest) on January 1, 2011. By means of unbundling two
separate companies are created: a (regulated) network company and a commercial energy company.
1.2

Effects of Evolution on Environment, Economics, and Security of Supply

The described evolution in the previous section had considerably effects for the three main domains of
this research work, environment, economics, and security of supply. These effects are described for
the central electricity production. The decentral electricity production i.e electricity geneation at private
owners, industry and distribution companies, waste incinerators, and renewable sources are also of
influence of the three mains domains. However, until the mid 90s these influences were limited and
therefore its influence on environment and economics are not described in detail.
1.2.1

Environment

Electricity production companies have always had a major influence on the environment. In fact, the
electricity sector was showing an interest in the environment long before pollution became a

Chapter 1 Introduction

mainstream political issue. It began with local problems arising from dust emitted by coal-fired power
plants. Attention has then shifted to regional problems (SO2 and NOx) and now focuses on the global
CO2 problem.
120

100

Index (1980 = 100)

80

CO2

100 = 35 Mton/a

NOx

100 = 84 kton/a

SO2

100 = 196 kton/a

E-Prod 100 = 58,100 GWh


60

Target SO2 (18 kton/a)


Target NOx (35 kton/a)

40

20

0
1980

1984

1988

1992

1996

2000

2004

Figure 1.6: Emission levels 1980-2002 of central electricity production (Source: RIVM, MNP)
Dust
Without technical measures coal- and oil fired power plants emit high levels of dust. These emissions
were significantly reduced in the 1970s through licensing requirements and new technology.
SO2
SO2 is emitted mainly by coal- and oil fired power plants. SO2 occurs when sulphur in the fuel
combines with oxygen from the combustion air. The subsequent compound is a major cause of acid
rain. SO2 emissions have been greatly reduced (see Figure 1.6) by licensing requirements, tighter
restrictions on sulphur levels in coal and oil (down from 2% to a maximum of 1%), and the use of flue
gas desulphurization systems.
NOx
Every time fossil fuel is burnt with oxygen from the combustion air, NOx is formed from nitrogen and
oxygen. NOx is harmful to the environment because it causes acid rain and smog. NOx emissions have
been strongly reduced (see Figure 1.6) and can be further reduced by licensing requirements coupled
with improved technology for gas turbines, burners in boilers, and DeNOx systems for coal-fired power
plants.
These three problems have clearly been brought under control by technological solutions which are
costly but are now mandatory via licensing systems and regulation. Therefore focus in this research,
concerning environment, is on the CO2 emissions.
CO2
Whenever fossil fuel (carbon combinations) is burnt with oxygen from the combustion air, CO2 is
released. It adds to the greenhouse effect. The CO2 emissions are strongly determined by the
composition of the primary fuel, the energy conversion method, the size of the demand, and the
approach to import and export. This problem is a long way from being solved. In fact, total CO2
emissions are currently rising (RIVM, 2004).

Fuel diversification
Figure 1.7 shows a clear switch from coal to gas towards the end of the 1960s, a return (to some
extent) to coal and oil in the early 1980s, and the return to gas and coal again. The composition of the
fuel package has a particularly strong effect on CO2 emissions.
650,000
600,000
550,000
500,000

Fuel consumption [TJ]

450,000
400,000
Biomass

350,000

Uranium
Gas

300,000

Oil
250,000

Coal

200,000
150,000
100,000
50,000

2004

2002

2000

1998

1996

1994

1992

1990

1988

1986

1984

1982

1980

1978

1976

1974

1972

1970

1968

1966

1964

1958

Figure 1.7: Fuel diversification in the Dutch central electricity production (Source: SEP, EnergieNed,
CBS)
Fuel Efficiency Delivered by Different Conversion Methods
The volume of primary fuel required per produced kWh of electricity halved in 1920 - 1950 and the fuel
consumption of a modern CCGT is now again half as much as the fuel consumption of a coal-fired
power plant in 1950. This reduction delivers huge cost savings as well as a decrease in environmental
pollution. The efficiencies in Figure 1.8 are till 1970 for coal/oil-fired power plants and after 1970 for
natural gas-fired power plants.
60%

50%

Net Electric Efficiency [%]

40%

30%

20%

10%
Average E - Efficiency coal/oil fired power plant
Maximum E - Efficiency gas fired power plant
0%
1920

1925

1930

1935

1940

1945

1950

1955

1960

1965

1970

1975

1980

1985

1990

1995

2000

2005

Figure 1.8: Development of net electric efficiency in full load operation (SEP, EnergieNed, Vlijm)

10

Chapter 1 Introduction

The key question at the moment is: How serious is the CO2 problem, especially in relation to the
climate, and how can it be solved at a global level? For, it is only at a global level that a solution can
be found! That makes CO2 emissions one of the greatest challenges facing us today. Meanwhile, the
next problem, the availability and distribution of energy sources, is already on the horizon.
1.2.2

Economics

After World War II the main priority was to re-establish an operational energy supply as soon as
possible. In the 1970s the emphasis rested on availability, raising the yield from fuel, and then, on
environmental considerations and the utilization of residual heat. All of these factors were instrumental
in creating the following situation:
Fuel Costs
At the end of 2003 fuel costs accounted for almost 70% of the total costs of the best available gasfired power plant, for coal-fired power plants this is rougly 50% (depending on actual fuel prices, raw
material prices and discounts factors). The fuel consumption per kWh has declined drastically (see
Figure 1.8) and the situation has been further improved by ever-increasing cogeneration. The sector
however has scarcely any influence on fuel prices, which tend to fluctuate widely over time depending
on the world politics (see Figure 1.9).
9.0
COAL / IMP COAL
OIL

8.0

NAT GAS

7.0

Fuel Price [/GJ]

6.0

5.0

4.0

3.0

2.0

1.0

2005

2003

2001

1999

1997

1995

1993

1991

1989

1987

1985

1983

1981

1979

1977

1975

1973

1971

1969

1967

1965

1963

1961

1959

1957

1955

1953

1951

1949

1947

1945

1943

1941

1939

1937

1935

1933

1931

1929

0.0

Figure 1.9: Average purchase value (money of the day) of fuel supply (Source: SEP, EnergieNed,
CBS, own information)
Investment Costs
Investment costs are the second-largest item (after fuel costs) of expenditure for coal- and gas-fired
power plants and the largest item of expenditure for nuclear power plants and sustainable production
options.
Other Costs, Especially Personnel Costs
The high share of fuel and investment in the overall production costs and the strong focus on security
of supply, amongst other things, meant that for a long time very little attention was paid to optimizing
personnel costs. Recently economics of scale have led to a strong reduction in the personnel costs
per product unit.

11

Taxes and prices


The electricity production and distribution companies have no influence on taxes. VAT and Energy Tax
(EB) account for around 50% of the total electricity rates for small-scale consumers. The electricity
prices were at lowest around 1940 (Vlijm, 2002), some peaks in the 70s and early 80s, nearly stable
till the mid 90s and strong rising since the mid 90s, first by rising taxes and the last few years by
rising oil- and gas prices.
1.2.3

Security of Supply

Security of supply implies that enough electricity production capacity is available to meet the
(fluctuating) electricity demand at all times3. It can be expressed as a percentage of the total time in
one year that demand can not be fulfilled or as a reserve factor which is: (available electricity
production capacity peak demand) / peak demand x 100%. From the end of World War II until the
mid-1970s everything had to be set in motion to meet the demand, the reserve factor was around 15%
or less. Thereafter, the growth in demand slumped and this figure rose from 15% to almost 30%. It
peaked at around 35% in 1995, whereupon it fell consistently to almost 0% in 2002 (see Figure 1.10)
with a heavy dependence on imported electricity from other countries. Before the liberalization of the
electricity sector, security of supply has always been approached from a predominantly technological
angle (100% production availability had to be guaranteed at all times) and costs have not been
weighed against the economic value of 100% security of supply.
Central installed capacity
Peak demand
64,000

14,000

56,000

12,000

48,000

10,000

40,000

8,000

32,000

6,000

24,000

4,000

16,000

2,000

8,000

0
1945
1948
1953
1958
1963
1964
1965
1966
1967
1968
1969
1970
1971
1972
1973
1974
1975
1976
1977
1978
1979
1980
1981
1982
1983
1984
1985
1986
1987
1988
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003

Net Production [GWh]

Installed capacity, Peak demand [MWe]

Electricity production
16,000

Figure 1.10: Development of the peakload and installed capacity of the Dutch central electricity
production (Source : SEP, TenneT, EnergieNed)
The ramp rate of all the central power plants combined is enough to handle the daily fluctuations in the
electricity demand in order to keep the equity between demand and supply (SEP, 1996). The influence
however of the decentralize power plants on the security of supply has increased starting at the early
90s specially caused by the turbulence grow after 1993 (see Figure 1.11) and the fact that most of the
decentralized power plants can not be ramped up or down as the central power plants.

Guaranteed fuel supply, security of supply of electricity transmission and distribution grids, and congestion
problems on the electricity interconnectors with Germany and Belgium are beyond the scope of this research.

12

Chapter 1 Introduction

8,000

Split-up 2005

7,500

IND CHP

7,000

414 64

1,312

CCGT DH
NON FOSSIL

6,000
Decentral installed capacity [MWe]

SMALL CHP

3,484

6,500

WASTE INC

1,795

247

OTHERS

5,500
5,000
4,500
4,000
3,500
3,000
2,500
2,000
1,500
1,000
500

2005

2004

2003

2002

2001

2000

1999

1998

1997

1996

1995

1994

1993

1992

1991

1990

1989

1988

1987

1986

1985

1984

1983

1982

1981

1980

1979

Figure 1.11: Development of the installed capacity of the Dutch decentral electricity production
(Source : SEP, TenneT, EnergieNed).
1.2.4

Observations

Is it necessary to steer a liberalized electricity market towards an optimal balance between


environment, economics and security of supply and, if so, what is the most desirable balance and how
far can this be realized? There is no simple answer to this question; first, because more experience is
needed, and second, because there are no national or international criteria for this balance. This
section will be limited to a few practical observations which will lead us into the research theme.
The main questions that need to be addressed in the current liberalized electricity market:
General
What is the most desirable balance between environment, economics and security of supply, and will
this be steered by the government or left to the market? The collective research task of the energy
companies has come to an end and the Annual Reports leave the impression that hardly any cash has
been set aside for this purpose.
The Environment
Will a liberalized electricity market address environmentally-friendly options such as sustainability,
cogeneration and district heating without additional regulation? The demand for energy continues to
grow sharply worldwide; meantime it is becoming increasingly clear that the environment and the
availability of gas and oil are (poised to become) a serious problem.
Economics
Many market players complain that the electricity market is not performing adequately because it lacks
transparency and there are still huge international disparities in legislation and regulations (Van Eck,
2004). This is prompting questions about the reliability of (international) price and market performance
comparisons. The situation is further compounded by the totally different starting positions of the
various countries with regard to the electricity production plants. As a result, it is almost impossible to
objectively compare a gas-driven producer such as the Netherlands with a nuclear energy-based
producer such as Belgium or France, a lignite/coal-based producer such as Germany, or a hydroelectricity producer such as Norway.

13

Security of Supply
Who is officially responsible for the long-term security of supply and is there enough transparency in
the overall situation? Is the economic climate conducive to investment in new and reserve capacity?
Many countries seem to be solving this via import along with the inevitable large inter-connector
capacity.
Chain Optimization
The split between production/trading/grids/retail/buyers does not lead to attention for optimizing the
whole chain of fuel supply/production/transport and distribution/buyers (chain dependence), despite
the fact that investments and operational decisions in parts of this chain have serious repercussions
for the chain as a whole. Further, there are no specific arrangements for district heating while the heat
market appears to be a natural monopoly.
Behavior of Market Players
The governments decision to liberalize the electricity market was a consequence of EU legislation and
was effectively based on the expectation that this would improve the electricity provision while
ensuring that the environment and security of supply were protected in the most efficient manner.
There was scarcely any discussion about the options, the choices that had to be made, and the best
way of steering such a process. The political parties also have totally different visions about how the
future electricity supply will pan out, ranging from an entirely free market dominated by conventional
large-scale electricity production to stringently regulated Distributed Generation based mainly on
sustainable electricity production options and energy-saving.
Figure 1.12 shows the entire electricity sector, consisting of the technological physical chain and the
politico-economic chain.
Physical Layer

Conventional Production
Nuclear
IGCC / CCGT
Renewables
Waste Incineration
Combined Heat & Power

Natural Gas
Oil
Coal
Uranium
Biomass
Waste
Others

Local Heat
Energy Intensive Industry
Utilities
Households

Distribution

Transport

Export

Import

Economic Layer

System Operations

Fuel Markets

Operationele dispatch

Commodity Market (OTC)


APX
Imbalance/reserve power
Import/export

Electricity Load

District heat
Heat Load
Proces heat

Law and Regulation

National Government

NMA / DTE

EU Directives

Figure 1.12: Schematic diagram of the electricity sector

14

NGO's
Knowledge Centres
Media

Chapter 1 Introduction

1.3

Research Objective and Questions

1.3.1

Research Objective

National and international research has been conducted in various domains, including technology,
economics, market operations, the environment, security of supply, regulation of public utilities and
liberalized markets, but no multidisciplinary research based on practice and theory has been carried
out to date. Observational data as discussed in the previous sections indicates that both the public
utility situation and the liberalized market do not automatically lead to an optimum for the balance
between environment, economics and security of supply. Supplementary measures to improve market
operations and/or regulations is required to achieve the goals for the balance as formulated by the
government.
The next section provides a brief survey of the main research projects on the operations of the
electricity sector. The graphical overview in Figure 1.12 shows all the areas of the electricity sector.
For the research questions of this study focus is on the physical and economic layer of the electricity
sector.
Overview of Relevant Studies
1.

Studies on the deployment of electricity production plants for a given trend in demand on the
basis of definable citeria:
Various models have been developed in this domain and are routinely applied by almost every
production owner in the free and liberalized market. The Dutch SEP electricity production plants
in the Public Utilities era was also deployed using an in-house adapted version of one of the first
PowrSym models. These models are generally ideal for optimal deployment on the basis of
marginal fuel costs, other variable costs and the technological parameters of individual
installations, but have the following limitations:

Heat delivery commitments are mostly not included in the optimization but are
programmed as obligatory.

Preconditions from the electricity grids are treated as obligatory and there is no
optimization effort for the whole energy provision chain.

The starting point is always 100% security of supply with no attention to the scope
(cost/benefit analysis) of the buyer.

They are not suitable for overall cost assessments and cross-chain optimization. Nor can
they effectively incorporate the increasingly complex fuel contracts, especially for natural
gas.

2.

Energy sector studies on an optimal long-term supply:


One example of this type of study is the SEP Electricity Plan (1996), which worked out the longterm electricity production and grid planning for trends in demand, new development and
decommissions. This plan needed the approval of the Ministry of Economic Affairs. Many
countries and regions adopted a similar strategy.
The main characteristics:
a. Growth in demand is an autonomous prediction with little or no scope for elasticity in the price.
b. The main focus is on the potential of the company itself. Scope for import/export and other
suppliers are often left out of the frame.
c. There is no environmental assessment or economic evaluation of the costs/benefits of security
of supply across the chain.

3.

Studies on the value of energy for buyers and the cost of savings/shifts in demand among buyers
compared with the savings realized by the producers (Stoft, 2002; IEA, 2002).

4.

Studies on policy-making processes:


a. Studies by Kema and ECN to determine the rates under the MEP environmental energy
scheme (ECN/Kema, 2005; ECN, 2004; ECN, 2005).
b. Studies by ECN, CPB and others into the long-term potential (ECN, 2005; CPB, 2005).

15

c. Studies by ECN and NRG on the consequences of termination or continuation operation of


Borssele nuclear power plant (ECN/NRG, 2005).
d. Various studies on the effect of CO2 emission trade (CPB, 2002; LEI, 2007).
5.

Grid studies.
These fall into three main categories:
a. Grid studies that take account of the (in)capabilities of production and buyers (Tennet, 2002;
TenneT, 2005; E.on Netz, 2003).
b. Grid studies geared to smart solutions whereby grids are not a constraint in the free market
(Kling, 2002).
c. Grid studies on the (in)capabilities of Distributed Generation compared with large-scale
production (Bach, 2004).

6.

Studies on the environmental costs of the energy provision across the chain (IVM, 1997; Espeel,
2002).

7.

Studies on market forces and price formation (Boisseleau, 2004; DTE, 2006; DTE, 2007;
NMA/DTE, 2007).

What these studies all have in common is that they almost invariably focus on parts of the physical,
economic or regulatory aspect of the energy supply chain and/or on a maximum of two of our three
domains of environment, economics, and security of supply.
1.3.2

Problem Definition

The issues pinpointed in this research raise the following questions:


Which scenarios are viable for the Dutch electricity supply system with a view to the trends in
demand/ supply, protectionism vs globalization, and the importance of the environment in a
liberalized electricity market with full competition? What effects do these have on the three
objectives of energy saving/reduction of CO2 emissions, improved economic efficiency and the
maintaining of a high level of security of supply? And what government measures are needed
to support them?
These three objectives are approached as follows in this research:
1.

16

The first objective, energy saving/CO2 emissions, represents the environmental aspect and is the
prime topic in the current debate on the environment. It is becoming increasingly evident that the
reserves of fossil fuel, oil and gas are finite and that more and more fossil fuel needs to be
imported from (currently) politically unstable countries to meet the national requirement. This
problem is compounded by the need to contain CO2 emissions in order to prevent (anticipated)
irreversible climatic changes. Hence, further sustainability in the electricity supply is a matter of
extreme urgency. Needless to say, environmental questions involve more than saving energy and
cutting CO2 emissions; the others are mentioned in Section 1.1 but are not discussed in detail,
first, because they are outside the scope of this research, and second, because, between 1970
and 1990, the government imposed extensive technological measures via licensing requirements
in order to contain other urgent threats to the environment (see also Figures 1.6-1.8). These
consisted of:
Mandatory filter installations in coal-fired power plants to restrict dust and soot emissions. Oil
plays scarcely any role nowadays.
Mandatory desulphurization systems in coal-fired power plants to restrict acidification and,
when necessary, the extraction of sulphur from natural gas for a few production sources.
DeNox installations, improved incineration technology and further improvements to the gas
turbine technology to restrict NOx in the exhaust fumes from coal- and gas-fired power plants.
Improved fuel efficiency.

Chapter 1 Introduction

2.

The economics aspect is economic efficiency. Cost-effectiveness and cost allocation are both
examined. Cost-effectiveness is a question of realizing goals in the most cost-effective way and
objective economic comparisons of various alternatives. Accurate cost allocation is vitally
important in the electricity sector because some parts of the chain can have far-reaching positive
or negative impacts on other parts. For example, large volumes of wind energy in the system
have serious implications for the infrastructure and the required reserve capacity of the remaining
production plants. This research particularly explores the economic aspects of the production and
the chain dependence. It does not explicitly address the economic aspects of the other parts of
the electricity supply chain.

3.

Security of supply is studied in relation to the technological/economic availability of electricity at


source. The availability of electricity transport or distribution infrastructure, fuel availability,
vulnerability to terrorist attacks, and the balance between the costs of security of supply and the
economic value of security of supply for the buyers is not studied.

To address the above-mentioned problem definition a technical-economic electricity production


simulation model addressing the technical system of the energy demand (electricity and heat) and
supply side is developed. Technology developments of the most important supply options based on
fossil fuel, renewables options, carbon capture and storage but also storage of electricity are included
and can be simulated. With use of this model four developed and possible future scenarios for the
Dutch electricity supply system are simulated. However, with the model all kind of scenarios can be
simulated because input parameters can be changed easily. The model results are based on a
transparent and controllable technical-economic world with a perfect working electricity market and no
impact from the physical chain of fuel supply, production, grids and consumers (chain dependence).
The implications of these external influences and the chain dependence which influences the
technical-economic system are described by Van Eck (2007).

History
and
research
definition/
thesis
Introduction

Reference
situation
(2003)

Part I

Technical
system
Transport
&
Distribution
Part II

Technical chain
- Fuel Supply
- Production
- Grids

Technical
developments
supply system

Scenarios and
Simulation
model

Part III

Part IV

New
Developments

Input model
and analyses
scenario
outcomes

Conclusions /
Recommendations

Extensive
Summary
for
policymakers

Part V

Model
- Scenarios
- Routines / Dispatch
- Demand / Supply
- Optimization

Scenario's
strategic
goals

Figure 1.13 : Total overview of this research

1.3.3

Research Questions

The research objective and problem definition will lead to the following subsidiary questions for this
study.
Part I

Description of the Reference Situation in 2003

How does the Dutch electricity sector perform in terms of environment, economics (Chapter 2), and
security of supply (Chapter 3) in the reference situation in 2003?

17

Part II Description of the System


What is the composition of the currently used supply side of the electricity and heat supply systems
and what are the technical conditions for unit deployment, environmental conditions and security of
supply (Chapter 4). What influences on the demand and supply side can be expected caused by
developments in electricity, gas and district heat grids (Chapter 5)?
Part III Supply Trends
In which way may the composition of the electricity supply side change in the future and what are the
technical developments which play an important role in the possible future composition of the
electricity supply side (Chapter 6).
Part IV Dutch Electricity Demand and Supply Scenarios and Simulation Model
Which scenarios for a future energy demand and supply system can be developed and what are the
most important driving forces and uncertainties (Chapter 7). How can a model be designed in order to
answer the questions for the balance between environment, economics and security of supply for the
developed scenarios and stated in the problem definition (Chapter 8).
Part V Input Model and Analysis of Scenario Outcomes
What data serves as input for the defined future scenarios (Chapter 9) and what are the differences
between the results for the calculated scenarios (Chapter 10).
1.3.4

Scope of the Research

This research is divided into five parts (Figure 1.13). The reference situation in the electricity sector is
described in Part 1 and is a starting point for this research. Part I presents an observation-based
description for an overall assessment of the performance of the electricity production sector in relation
to the environment, economics (Chapter 2), and security of supply (Chapter 3) within the contours of
the current legislation and regulations and market structure.
Part 2 describes the current electricity supply system, divided in conventional and combined cycle-,
combined heat and power- and sustainable production options in the electricity sector. These are
analyzed (Chapter 4) and their technical conditions for supply, the environmental situation and
influences on security of supply are described. In combination with the described transport and
distribution of energy (Chapter 5) it shows how the various parts of the chain are connected.
In Part 3 the technological developments for electricity production, divided also in conventional and
combined cycle-, combined heat and power- and sustainable production options, CO2 capture and
storage and electricity storage possibilities are explained and worked out resulting in an overview
matrix (Chapter 6) with all possible options and parameters (power ranges, efficiencies, operations
and maintenance cost and investment cost).
Part 4 of the research begin with scenario development for possible future energy scenarios (Chapter
7) and four possible scenarios are formed. The mathematical/economic structure of the simulation
model, the departure points, the mathematical algorithms, the required input, and the generated output
are expained (Chapter 8) and the model is validated.
In Part 5 the demand side is analyzed and quantitatively defined as input for the model while the
supply side also addresses the role of the various production options in covering the demand (Chapter
9) for mid- and long term development as function of the applicable scenario. The graphical outcomes
of the model calaculations for the scenarios are presented. The differences between the scenarios
are analyzed and summarized (Chapter 10).

18

Chapter 1 Introduction

In Conclusions and Recommendations the results from the different parts of this study are used to
further underpin the answers to the research questions. This leads to quantitative recommendations
based on results from the model and substantiated recommendations about choices which can be
made for possible future scenario and further investigation with use of the simulation model.
1.4

Research Overview

Introduction
Chapter 1 :

Introduction

Part 1 : Description of the Reference Situation in 2003


Chapter 2 :
Environment and Economics
Chapter 3 :
Security of Supply
Part 2 : Description of the System
Chapter 4 :
Energy Supply Chain and Supply Options
Chapter 5 :
Transport and Distribution of Energy
Part 3 : Supply Trends
Chapter 6 :
Technological Trends in Supply Options
Part 4 : Dutch Electricity Demand and Supply Scenarios and Simulation Model
Chapter 7 :
Scenarios for the Dutch Electricity Supply
Chapter 8 :
Dutch Electricity Supply Model
Part 5 : Input Model and Analysis of Scenario Outcomes
Chapter 9 :
Input Model for the Developed Scenarios
Chapter 10 :
Analysis of Scenario Outcomes
Conclusions and Recommendations
Chapter 11:
Overall Conclusions

19

References

References
Algemene Energieraad (2005), Gas voor morgen [Gas for Tomorrow]
Algemene Energieraad (1978), Energiebesparing in bedrijven [Saving Energy in Businesses]
Bach, P-F (2004), Impact of Distributed Generation on System Operation, Eltra, the TSO of the
western part of Denmark
Boisseleau, F. (2004), The Role of Power Exchanges for the Creation of a Single European Electricity
Market: market design and market regulation, DUP Science, ISBN 90-407-2466-0
Commissie Concentratie Nutsbedrijven (1978), Coconut rapport en regeringsstandpunt mbt
schaalvergroting bij elektriciteitsproductie concentratie en horizontale integratie bij energiedistributie
[Coconut Report and Government Position on Economies of Scale in the Concentration of Electricity
Production and Horizontal Integration in Energy Distribution]
CPB (2002), Economische effecten van nationale systemen van CO2-emissiehandel, ISBN 90-5833088-5 [Economic Effects of National Systems of CO2 Emission Trading]
CPB (2005), Werkgelegenheid en toegevoegde waarde per bedrijfstak, 2001-2020 en 2021-2040,
CPB Memorandum [Employment and Added Value per Business Sector, 2001-2020 and 2021-2040]
DTE (2006), Marktmonitor, ontwikkeling van de groothandelsmarkt voor elektriciteit in 2005,
Projectnummer: 102214 [Market Monitor: development of the wholesale electricity market in 2005]
DTE (2007), Marktmonitor, ontwikkeling van de groothandelsmarkt voor elektriciteit in 2006,
Projectnummer: 102641 [Market Monitor: development of the wholesale electricity market in 2006]
ECN/Kema (2005), Inzet van biomassa in zelfstandige kleinschalige installaties voor de opwekking
van electriciteit, Berekening van de onrendabele top, ECN-C--05-016 [Deployment of Biomass in
Stand-Alone Small-Scale Electricity Generation Plants: Calculation of the Unprofitable Top]
ECN/Kema (2005), Inzet van biomassa in centrales voor de opwekking van electriciteit, Berekening
van de onrendabele top, Eindrapport, ECN-C--05-088 [Deployment of Biomass in Electricity
Generation Plants: Calculation of the Unprofitable Top]
ECN (2004), Advies WKK MEP-tarief 2004, ECN-C--04-049 [Recommendations CHP MEP Rate 2004]
ECN (2004), Advies WKK MEP vergoeding 2005, ECN-C--04-111 [Recommendations CHP MEP
Compensation 2005]
ECN (2005), Advies WKK MEP vergoeding 2005, ECN-C--04-111a [Recommendations CHP MEP
Compensation 2005]
ECN (2005), MEP-advies WKK 2006, ECN-C--05-102 [MEP Recommendations CHP 2006]
ECN (2005), Referentieramingen energie en emissies 20052020
Projections for Energy and Emissions 20052020].

ECN-C--05-018 [Reference

ECN (2005), The next 50 years: Four European Energy Futures, ECN-C--05-057.
ECN/NRG (2005), Kerncentrale Borssele na 2013, Gevolgen van beeindiging of voortzetting van de
bedrijfsvoering, ECN-C--05-094 [Nuclear Power Plant Borssele, Consequences of termination or
continuation operation]
EnergieNed (2000), Kerngegevens ten behoeve van de Energiesector 2000 [Core Figures for the
Energy Sector 2000]

20

References

Espeel, E.J.H. (2002), The EECE Method, EV-2105, Thesis Delft University of Technology
E.on Netz (2003), Wind Power in Germany, Present situation and outlook, Presentation January 23 rd
2003
European Commission (1996), Directive 96/92/EC of the European Parliament and of the Council of
19 December 1996 concerning common rules for the internal market in electricity
European Commission (1998), Directive 98/30/EC of the European Parliament and of the Council of
22 June 1998 concerning common rules for the internal market in electricity
European Commission (2003), Directive 2003/54//EC of the European Parliament and of the Council
of 26 June 2003 concerning common rules for the internal market in electricity and repealing Directive
96/92/EC
Groenewegen, J. et al. (2004), An Evolutionary Perspective on Infrastructure Reform: the case of the
Dutch electricity sector, paper for the EAPE conference, Crete Oct. 2004.
Hesselmans, A.N. (1996), Wisselende spanning, een historische verkenning naar de relatie tussen
Rijksoverheid en elektriciteitssector [Alternating currents, a historical survey of the relation between
government and electricity sector]
IEA (2002), Distributed Generation in Liberalized Electricity Markets, ISBN 92-64-19802-4
Institute for Environmental Studies (IVM), VU University Amsterdam (1997), External National
Implementation the Netherlands, Contract Jos3-CT95-0010
Kling, W.L. (2002), Intelligentie in netten: modekreet of uitdaging, Intreerede TU Eindhoven, 7 June
2002 [Intelligence in Grids: fashion statement or challenge?]
Kper, N. (2003), Tegenpolen, de liberalisering van de Nederlandse energiemarkt, ISBN 9027487367
[Opposite Poles: the deregulation of the Dutch energy market]
Knneke, R.W. (1996), De rol van lokale en regionale overheden in een geliberaliseerde
elektriciteitsmarkt [The Role of Local and Regional Governments in a Free Electricity Market]
LEI (2007), Emissiehandel voor glastuinbouw : Effecten van een CO2-vereveningssysteem, ISBN
9789086151646 [Emissions Trading for the Market Garden Sector: effects of a CO2 equalization
system]
Ministry of Economic Affairs, the Netherlands (1995), Derde Energienota [Third Energy Policy
Document]
Ministry of Economic Affairs, the Netherlands (2000), Energie en samenleving in 2050 [Energy and
Society in 2050]
NMA / DTE (2007), Gasmonitor, Ontwikkeling in de groothandelsmarkt gas in Nederland in 2006,
Projectnummer: 102642 [Gas Monitor: developments in the wholesale gas market in the Netherlands
in 2006]
PGEM (1988), Huishoudelijke apparatuur, Informatie Nachtstroom, Brochure [Household Appliances,
Information on Night Power]
RIVM (2004), Milieubalans 2004, ISBN 90-6960-109-5 [Environmental Balance 2004]
SEP, GKN, KEMA, VEEN, VDEN (1972), Elektriciteit in Nederland 1972 [Electricity in the Netherlands
1972]
SEP, GKN, KEMA, VEEN, VDEN (1972), Elektriciteit in Nederland 1983 [Electricity in the Netherlands
1983]

21

References

SEP, GKN, KEMA, VEEN, VDEN (1972), Elektriciteit in Nederland 1984 [Electricity in the Netherlands
1984]
SEP, EnergieNed (1998), Elektriciteit in Nederland 1998 [Electricity in the Netherlands 1998]
SEP (1996), Elektriciteitsplan 1997 2006 [Electricity Plan 1997 2006]
SEP (1996), Toelichting bij het Elektriciteitsplan 1997 2006 [Notes to the Electricity Plan 1997-2006]
Staten-Generaal der Nederlanden (1998), Wet van 2 juli 1998, houdende regels met betrekking tot de
productie, het transport en de levering van elektriciteit (Elektriciteitswet 1998) [Electricity Act 1998]
Staten-Generaal der Nederlanden (1989), Elektriciteitswet 1989) [Electricity Act 1989]
Stichting Historie der Techniek (2000), Techniek in Nederland in de twintigste eeuw, ISBN Deel II 905730-064-8 [Technology in 20th-Century Netherlands]
Stoft, S. (2002), Power System Economics, Wiley-IEEE Press, ISBN 0-471-15040-1
TenneT (2002), Capaciteitsplan 2003-2009 [Capacity Plan 2003-2009]
TenneT (2005), Kwaliteits-en Capaciteitsplan 2006-2012 [Quality and Capacity Plan 2006-2012]
Van Eck (2007), A New Balance for the Energy Sector: no longer a puppet in the hands of technology,
public interests and market, ISBN 978-90 788 8901-4
Vlijm, W. (2002), De interactie tussen de overheid en de elektriciteitssector in Nederland, ISBN 909015850-2 [Interaction between the Government and the Electricity Sector in the Netherlands]
Vringer, K. (2005), Analysis of the Energy Requirement for Household Consumption, Milieu en Natuur
Planbureau, Bilthoven. Thesis, ISBN: 90-6960-130-3

22

Part 1

Description of the Reference Situation in 2003

24

Chapter 2 Environment and Economics

Chapter 2
2

Environment and Economics


2.1

Introduction

Part I presents an observation-based description of the performance of the energy production sector
for the reference situation in 2003. The first part of Chapter 2 covers the environmental (CO2)
performance of the energy production sector. The second part is focusing on the economic aspect of
electricity production versus market price trends for peak and off peak periods in 2003. In Chapter 3
the balance between electricity demand and generation capacity is described and the security of
supply is analyzed.
2.2

Environmental (CO2) Performance

The latter years of the twentieth century saw a considerable improvement in the environmental
performance of Dutch power plants (see Introduction), particularly to CO2 emissions as explained in
this Chapter. This was largely due to the deployment of new technologies to improve efficiency and
reduce emissions, the decommissioning of obsolete units, the method of operation (especially the
National Economic Optimization (LEO) in the context of the Association of Energy Producers (SEP)),
legislation, and the increase in CHP and Renewable potential. Nuclear energy is a separate issue.
Nuclear power stations and Renewables deliver equal performances in reducing CO2 emissions. At
present, nuclear energy is again a topic of political/social debate. The closure of the only nuclear
power plant in the Netherlands has now been postponed to 2033.
In 1998 the Netherlands signed the Kyoto Protocol along with the other EU member states. Climate
negotiations in Bonn and Marrakech ended with international agreements on the implementation of
Kyoto. It was agreed that the Netherlands would contribute to Europes commitments under this
protocol by cutting greenhouse gas emissions by an average of 6% in 2008 - 2012 compared with
1990.
In 1999 and 2000, both parts of the Memorandum on the Implementation of Climate Policy
(Uitvoeringsnota Klimaatbeleid (UK)) explained how the Netherlands would meet its Kyoto
commitment. In 1990 the level of greenhouse gas emissions in the Netherlands was approximately
212 Mtons (see Figure 2.1). To reach the Kyoto target of 6% reduction, an average emission level of
199 Mtons per year needs to be achieved in the first phase (2008 - 2012). In the business-as-usual
scenario (without the UK Memorandum) it is estimated that greenhouse gas emissions would rise to
approximately 239 Mtons of CO2 equivalents. Bearing in mind the policy outlined in the UK
Memorandum, the total reduction attained in 2010 by the UK measures would be 14 Mtons CO2 eq.,
resulting in greenhouse gas emissions of 225 Mtons CO2 eq.
This means that, even when the UK measures have been implemented, there is still a shortfall of 26
Mtons CO2 eq. According to the Dutch government 50% of the total reduction target can be achieved
through domestic efforts; the rest will have to come from elsewhere.
Some of the UK measures for the reduction of 14 Mtons CO2 eq. will have to be implemented by
industry, including the electricity production companies. The Dutch governments focus is on an
agreement on coal with the owners of coal fired power plants and an agreement on energy-efficiency
benchmarking with the electricity producers and the major energy consumers. The electricity
production companies in the Netherlands have agreed on a collective annual reduction of 6 Mtons
CO2 eq. by 2012.
With the implementation of the (European) Emission Trading System (ETS) in 2004, CO2 emissions
have a market value. The CO2 price is part of the electricity cost price and clearly, all consumers have
to pay this price. With use of the national allocation greenhouse gas emissions for energy sector and

25

Part 1

the industries (first allocation period 2005 - 2007, second 2008 - 2012) the Dutch government (in
dialogue with EU) can influence the total reduction target by means of tightness in allocation rights.

Figure 2.1: Effects of UK and Kyoto objective (Source: Ministry of Housing, Spatial Planning and the
Environment, 2002)

2.2.1

CO2 Emissions Electricity Production Sector

The production of electricity and heat of the electricity production companies and other electricity
producers in the Netherlands leads to CO2 emission. For both the CO2 amount will be given /
calculated for 2003. The CO2 emissions of the imported electricity do not count for the total in the
Netherlands, but a calculation is made to get a feeling about the amount of this CO2 emissions. The
total amount of the CO2 emissions of the electricity production sector and import is used to have a
good reference for comparing the future scenarios calculated with the model of the Dutch electricity
supply.
Electricity production companies: for a net electricity production of 64.0 TWh (see Table 3.1) the
produced CO2 amounts in total 42.4 Mton (Environmental Reports Production Locations, 2003) in
2002 (as check); this was 40.2 Mton for a net electricity production of 63 TWh (Milieu &
NatuurCompendium, 2004).
Other electricity generation: for a net electricity production of 28.8 TWh1 (see Table 3.1) of which 1.9
TWh non fossil (1.6 TWh non fossil + 0.3 TWh gas engines biogas) and 2.0 TWh Waste Incineration,
the produced CO2 amounts in total 21.1 Mton. The calculation of the CO2 emissions of CHP and small
scale CHP is based on the fuel consumptions figures given by CBS table with separate electricity
production divided in clusters. The amounts of CO2 for waste incineration is calculated based on a
methodology presented in Netherlands National Inventory Report 2005 of RIVM (2005)2.
1

The net electricity production of 28.8 TWh (gross 30.2 TWh) of other electricity production is based on table
Electricity Balance of CBS Statline. The calculated CO2 emissions is based on the productions used in CBS
Table with separate electricity production divided in clusters. This electricity production is here gross 31.6 TWh
and net 30.2 TWh (calculated) and thus 1.4 TWh higher. According to CBS (2006) the last is correct.
2

For waste incinerations 47% of the applied waste is non fossil (situation 2003) and therefore no CO2 emissions
of fossil origin are accounting for this part.

26

Chapter 2 Environment and Economics

Import: a total of 17.0 TWh (net) was imported in 2003 of which 9.7 TWh was non fossil and imported
via certificates (CBS, 2005). No information was found if this non fossil generated electricity was extra
generated above the normal production in the exporting country or that the exported non fossil
electricity had to be replaced by conventional electricity in the exporting country and was mainly
caused by differences in level playing field between EU countries (van Eck et al, 2004).
In the first case this would mean a saving on global CO2 emissions; in the second case this would
mean less savings on global CO2 emissions due to the composition of the installed production
capacity in the neighbouring countries. For this research, for reasons mentioned above, the second
case is assumed, so CO2 emissions are accounted for the 9.7 TWh imported non fossil electricity.
According to Van Eck et al (2004) France and mainly Germany are the major exporting countries to
the Netherlands. Looking to the physical electricity flows in 2003 (UCTE), the import from Germany
was 85% and from France 15% of the total import, see Annex A. For Germany, with a share of
approximately 27% nuclear and 50% lignite and coal of the gross produced electricity, a CO2 factor
(based on the net electricity production) of 0.6 Mton/TWh is calculated. For France, with a share of
almost 80% nuclear, a CO2 factor (based on the net production) of 0.06 Mton/TWh is calculated. With
the assumptions made concerning the imported non fossil electricity combined with the physical
electricity flows in 2003 an amount of 8.8 Mton CO2 emissions is calculated for imported electricity.
The total CO2 emissions in 2003 caused by the production of electricity and heat in the Netherlands
therefore amounts to: 63.5 Mton. Including the CO2 emissions of imported electricity this amounts to:
72.3 Mton.
2.2.2

Effects of Choices in Electricity Production Options on CO2 Emissions

The examples below help to explain the effects on CO2 emissions of choices made in electricity
production options, total performance of the energy supply system and re-use of waste heat (based
solely on direct emissions):
In 2003, approximately 1.9 TWh (see Table 9.9 of Chapter 9) of renewable electricity was
generated annually in the Netherlands (excluding bio-energy from waste incinerators and coalfired power plants with partial biomass firing). If this were to replace gas-generated electricity
with an average efficiency of 50% it would cut annual emissions by approximately 0.8 Mtons.
If it were to replace CHP capacity (with a 20% saving on fuel compared with the abovementioned gas capacity), the annual reduction in emissions would be approximately
0.6 Mtons.
Each percentage point in the reduction of efficiency across the whole spectrum of electricity
generation (e.g. from 44% to 43%) which takes place as a result of sub-optimization by each
market player increases emissions by approximately 1.5 Mtons a year. These calculations are
based on the annual generation rate of 110 TWh (2003) and an average gas/coal ratio of
70/30.
If half the residual heat from a modern gas-fired 350 MW CCGT (less than 2% of Dutch
installed capacity) with an operating time of 6,000 hours were used for heating purposes, it
would deliver annual emission savings of approximately 0.2 Mtons.
Shutting down the Borssele nuclear power station and replacing it with modern gas capacity
would produce additional annual CO2 emissions of approximately 1.4 Mtons and, if replaced
with coal-fired capacity, approximately 3.1 Mtons.
2.3

Economic Performance

In the free electricity market prices are based on short- and long-term agreements between sellers
and buyers. The prices on the daily spot market can fluctuate widely because of short-term capacity
problems. Stabilization is discernible in longer-term agreements through the emergence of two prices:
one for off-peak periods (the whole weekend, holidays and working days from 23.00 - 07.00 hours)
and one for peak periods (working days from 07.00 - 23.00 hours). Our analysis was conducted for
these two prices alone for the reference situation in 2003 and is based on the available capacity and
the average cost parameters of the different electricity production units.

27

Part 1

2.3.1

Electricity Market Pricing in 2003

The reference situation of 2003 is characterized by a large number of suppliers, including around 20%
import and the following electricity markets:

Peak 03

Base 03

Off Peak 03

Gas 03

27-12-02

12-12-02

27-11-02

12-11-02

28-10-02

13-10-02

28-09-02

13-09-02

29-08-02

14-08-02

30-07-02

0
15-07-02

0
30-06-02

10

15-06-02

10

31-05-02

20

16-05-02

20

1-05-02

30

16-04-02

30

1-04-02

40

17-03-02

40

2-03-02

50

15-02-02

50

31-01-02

60

16-01-02

60

1-01-02

Electricity price [/MWh]

Coal 03

Figure 2.2: Forward electricity and fuel price in 2002 for 2003 (Source: Endex, EnergieNed, own
information)

28

Gas price [ct/m3], Coal price [$/Mton]

A commodity market for domestic and foreign purchases and sales based mainly on forward
(usually annual) contracts with an average price for 2003 of 31.4/MWh for base load (8760
h), 47.6/MWh for peak load (workdays 7:00-23:00, 4064h) and 17.4/MWh for off peak load
(nights 23:00 7:00, weekends and holidays, 4696 h); see Figure 2.2. This market constitutes
85 - 90% of the total market.
A commodity spot market. The remaining 10 -15% is traded through daily offers (day-ahead)
via the APX. Price movements on the APX in 2003 were somewhat unpredictable (see
Figures 2.3 and 2.4). The extreme peaks in prices coincide with disruptions in the market
caused by, for example, the unexpected outage of units, a restriction on import capacity, or a
shortage of cooling water caused by high river water temperatures on summer days.
A market for control and reserve capacity. TenneT is the buyer and each electricity producer
may be a seller.
A market for import/export capacity. This market is organized by TenneT and regulated by
DTe.

Chapter 2 Environment and Economics

250

Electricity price [/MWh]

200

150

100

50

0
Monday, 29
Tuesday, 30
September, 2003 September, 2003

Wednesday, 01
October, 2003

Thursday, 02
October, 2003

Friday, 03
October, 2003

Saturday, 04
October, 2003

Sunday, 05
October, 2003

Monday, 06
October, 2003

Figure 2.3: Price trends on the APX in week 40, 2003 (Source: APX)

2,250

2,000

1,750

Electricity price [/MWh]

1,500

1,250

1,000

750

500

250

0
Monday, August Tuesday, August
11, 2003
12, 2003

Wednesday,
Thursday, August
August 13, 2003
14, 2003

Friday, August
15, 2003

Saturday, August Sunday, August


16, 2003
17, 2003

Monday, August
18, 2003

Figure 2.4: Price trends on the APX in week 33, 2003 (Source: APX)

As these are commodity (MWh) markets, they exclude transport costs and end-user taxes.

29

Part 1

2.3.2

Anticipated Trends in Peak and Off-Peak Prices for 2003

The analysis was performed in three steps based on the following cost components of the electricity
production:
1. The fuel cost per MWh.
2. The dispatch plus fixed exploitation costs per MWh. Dispatch cost are fuel costs plus variable
operation and maintenance cost. Fixed exploitation costs are additional capacity costs for
natural gas and fixed operation and maintenance.
3. The total or integral cost or per MWh. Integral cost are dispatch plus fixed exploitation cost
plus the investment cost (depreciation, interest). The investment costs were based on average
key figures for each type of capacity around the year 2003. For cost of financing a 15 year
annuity at an interest rate of 6% was applied.
The available generation capacity was divided into four clusters:
1. Conventional coal.
2. Combined Cycle Gas Turbine (CCGTs).
3. Hot Windbox CCGTs.
4. Conventional gas.
Average costs were established for each cluster. Fuel prices (commodity) 1.5/GJ (including the cost
of residue and ancillary materials) was used for coal and 3.8/GJ for gas. These are the average
forward prices in 2002 for 2003 (see Figure 2.2). The price of coal was fairly constant till 2004, then
first rose and fell again in 2006. The gas price fluctuates strongly; its effects have been investigated
but are not incorporated in this summary. Between the end of 2003 and the end of 2006 the price of
gas rose strongly (by around 100%).
The electricity market will not operate according to the known supply/demand models with price
elasticity, especially in weekday hours. This is because stock build-up is not possible as there is no
direct relationship between gas and coal prices and the technical possibilities and the costs of the
various conversion techniques are totally different. Any price elasticity in the demand is assumed to be
incorporated into the demand curve and the supply side therefore determines the pricing.
Figure 2.5 shows fuel costs as a function of operating time per cluster of generation capacity.
50

Fuel cost [/MWh]

40

30

CONV
HW CCGT
20

CCGT
COAL

10

0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

Operating time [h]

Figure 2.5: Fuel costs as a function of operating time at a coal price of 1.5/GJ and a gas commodity
price of 3.8/GJ

30

Chapter 2 Environment and Economics

As an overcapacity can always be expected during night and weekend hours, pricing in these periods
( 17.4/MWh for off peak load) will be based mainly on the fuel costs. In the import situation there is
certainly enough coal capacity available to meet the whole demand, not to mention mandatory
deployment (CHP with process steam and hot water supply without buffering, blast furnace gas,
Borssele nuclear power plant, waste incineration plants, grid commitments, control capacity).
When the total energy saving benefit of CHP is allocated to electricity, the fuel costs per kWh for good
CHP is approximately 20 - 25% lower than for modern CCGTs. When the gas price rises above
approximately 4.0/GJ, the fuel costs of the CHP units, including the MEP rebate for second half
2003, exceed the market price during nights and weekends.
Figure 2.6 shows the dispatch + fixed exploitation cost as a function of operating time per cluster of
generation capacity. As the units get almost no compensation for their non-fuel costs during nights
and weekends, they have to earn them back during daytime hours. The total number of daytime hours
is approximately 4,100. An operating time of between 3,500 and 4,000 full load hours is feasible for
coal during daytime hours, between 3,000 and 3,500 hours for CCGTs, and between 750 and 2,000
hours for the other clusters. Only units with low financial burdens or none at all can afford to compete
at this price level.
80

70

Dispatch + fixed cost [/MWh]

60

50
CONV
HW CCGT

40

CCGT
COAL
30

20

10

0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

Operating time [h]

Figure 2.6: Dispatch + fixed exploitation cost as a function of operating time at a coal price of 1.5/GJ
and a gas commodity price of 3.8/GJ
Figure 2.7 shows the total costs as a function of operating time per cluster of generation capacity.
Clearly, units with high financial burdens cannot afford to compete these price levels of approximately
47.6/MWh for peak load. They receive no compensation for their financial costs. These price levels
also reduce the likelihood of new capacity being built in the short term.

31

Part 1

110
100
90

Integral cost [/MWh]

80
70
CONV

60

HW CCGT
CCGT

50

COAL

40
30
20
10
0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

Operating time [h]

Figure 2.7: Total costs as a function of operating time at a coal price of 1.5/GJ and a gas commodity
price of 3.8/GJ
The electricity price will probably remain low during nights and weekends at a level equal to or slightly
above the cost of fuel for coal-fired units. This level for 2003 lies between 15/MWh and 20/MWh,
which, given the gas price for 2003, is even lower than the fuel cost for modern CCGTs operating at
base load. If gas prices rise steeply and there is enough import capacity, this increase will have only a
very limited effect on the off peak electricity price.
2.4

Summary

The environmental (CO2) performance of the energy production sector is analyzed and total CO2
emissions for the reference situation in 2003 are calculated. The economic aspect of electricity
production versus market price trends for peak and off peak periods are analyzed for four production
options3.

After 2003 the fuel - and electricity prices have raise strongly (see Figure 9.24 of Chapter 9) and as of 2004 the
Emission Trading System came into operation. For the comparison of the average costs components for each of
the 4 clusters, however, this means that the absolute cost levels have raise also but the relative differences
between the clusters did not change because the gas/coal price ratio did not change significantly.

32

Chapter 3 Security of Supply

Chapter 3
3

Security of Supply
In the Netherlands it is imperative to maintain a balance between supply and demand in electricity, as
large-scale storage is technically and economically out of the question in the Dutch situation. The
national electricity requirements must be met by generators inside the Netherlands and by means of
imported electricity. In the past, SEP compiled regular forecasts of the growth in demand for electricity
in consultation with user groups. As the central organization, SEP used these forecasts to work out the
program for building and decommissioning generating facilities, one of the criteria being maximum
avoidance of cases of inability to supply. In this system SEP was responsible for security of supply.
With the advent of the liberalized market and the end of the OVS, responsibility for security of supply
passed to the market players including the buyers on January 1, 2001.
In this chapter we concentrate solely on the relationship between supply and demand and provide an
indication of the risks based on a static approach. This will be explored further in Part V with the aid of
the model and chosen scenarios. Nor shall we address the potential of Demand Side Management
(DSM) to improve security of supply. This chapter focuses exclusively on the balance between
demand and generation capacity in the Netherlands and how this affects security of supply for the
reference situation in 2003.
3.1

Current Method of Operation

In the current set-up TenneT is the operator of the national grid and, as the Transmission System
Operator (TSO), it monitors the energy balance of the Dutch system as a whole. Transactions by
generators, suppliers, traders, brokers and consumers are recorded in E programs, which must be
reported to TenneT by the next day with details of quantity, time schedules, and transmitters and
recipients. Deviations from programs result in electricity transactions with TenneT within its Program
Responsibility. Deviations are dealt with by the so-called primary and secondary control systems.
The System Code denotes the UCTE rules for these control systems.
Primary control: The aim of the primary control system is to immediately restore as much as possible
of the balance between the required and supplied capacity at the generating units if a sudden failure
occurs in the interconnected European grid. In this context as much as possible means the
maximum reserve that needs to be activated by the primary control system (= primary reserve). This
amounts to 3,000 MW for all the UCTE partners combined. Hence, the primary control system should
be capable of handling a failure of a maximum of 3,000 MW.
The Dutch share in the UCTE primary reserve for 2003 (determined in proportion to its share of the
total volume of UCTE generation) was set at 120 MW (TenneT, 2002).
Secondary control: The purpose of the secondary control system is to ensure that, after a failure, the
power exchange with the other countries is returned to the desired level and the primary reserve again
becomes available by deploying control capacity in the country where the failure occurred. The
secondary control system is also used continuously in the Netherlands to smoothen any imbalances
that occur because of deviations from the E programs.
The control capacity is provided by suppliers who have signed a contract with TenneT and suppliers
who offer capacity on a voluntary basis. Suppliers provide TenneT with reserve capacity in addition to
control capacity.
3.2

Methods Applied to Analyze Security of Supply

Security of supply can relate to electricity production, grids, or a combination of the two. This research
deals only with production. The following methods have been applied so far to determine the risk of
non-guaranteed security of supply in the Netherlands:

33

Part 1

34

Method to determine the minimum required production capacity:


This method, applied by SEP (1996) in the Public Utility period, works as follows: the
maximum load for a period is defined as a probability density function with a normal
distribution. The forecast maximum annual load and the subsequent maximum loads for each
of the periods are anticipated values and the uncertainties (coincidences) are expressed as
the standard deviation. Starting with the anticipated probability of the non-availability of all
means of production, a function is defined for the probability that a specific demand cannot be
met by the production capacity that in a given period is unavailable (unplanned). The eventual
probability of non-capacity is determined by the product of the probability density for the
maximum load and the probability of non-capacity for a specific load.
The following criteria are applied:
o The probability of non-capacity must not exceed 2% during the peak load in any of the
periods.
o The geometric average of the probability of non-capacity during the peak load of the
13 four-week periods must not exceed 1% in a planning year.
o The probability of non-capacity must not exceed 1% in periods where there is little or
no revision during peak loads.
Electricity imports with capacity guarantees lower the minimum required capacity in the home
country. The ratio between the minimum required capacity and the maximum annual load is
the reserve factor.

The Dutch government has asked TenneT to monitor the development of supply and demand.
TenneT (2002) has devised a new assessment method for this purpose. The development of
this method was driven by the liberalized market with producers operating in it and no central
control, and no insight into the development of production capacity. Mainstream approaches to
security of supply, such as the reserve factor (see method for ascertaining the minimum
required production capacity), are less suitable because they are based on centralized control.
The new method works out how much production capacity is provided by the market players
and how it relates to trends in the electricity demand. It indicates how much elasticity i.e.
free space in production capacity exists in the electricity market to guarantee the balance
between supply and demand. The elasticity is assessed on the basis of two factors: the
anticipated production capacity in the coming years (producers estimates) and a demand
forecast compiled by the CPB (Bureau for Economic Policy Analysis). TenneT uses these
sources to calculate how much free space exists in the system so that sudden disruptions of
supply or fluctuations in demand can be evened out immediately. This method shows the
extent to which the Netherlands would depend on reserve capacity and/or imports to realize
the required free space in a future situation.
Advantages: TenneT (2002) names several advantages, but the most important is that the
model is based on observations and verifiable suppositions or hypotheses. The model shows
the expected/actual use factor for commodity capacity, which is a key consideration in
investment decisions. Limitations: The model does not give the absolute probability of breaks
in supply; it only shows the circumstances in which there are insufficient dispatch possibilities
to meet market obligations and commodity capacity.

To bolster this method a decision was taken to use an alternative framework based on the
LOLE approach (Beune et al, 2004). LOLE stands for Loss of Load Expectation and is used
worldwide as a system-adequacy indicator and for performing analyses in conjunction with
other countries in linked systems. It is based on the numerical mathematical determination of
the probability that the available production capacity will be too low to meet the overall
demand. LOLE models can deliver anticipated values for the amount of time that the
production capacity will not be capable of meeting the demand.
Input data for the demand side:
o Chronological load pattern of the total demand for all the hours in the calculation
period.
o Load uncertainty (standard deviation).
Input data for the supply side:
o Maximum production capacity.
o Percentage of planned non-availability (or a planned outage).

Chapter 3 Security of Supply

Probability of unplanned non-availability.

Essentially, this is the same method applied by the SEP.


3.3

Demand for Electricity

The demand for electricity in the Netherlands is divided between four sectors: industry, transport,
tertiary and residential consumption. Each sector has its own characteristic pattern of demand.
Economic growth and the Gross Domestic Product (GDP) are important factors when estimating
annual growth in consumption. Until 1999, the increase in electricity consumption always outstripped
the increase in GDP. In 1999 and 2000 the situation was reversed, but in 2001, 2002 and 2003 the
increase in electricity consumption was again (much) higher than the increase in GDP.
An electricity balance sheet was drawn up of electricity consumption and the various generation
categories. Table 3.1 shows this balance sheet for 2002 and 2003 (see appendix B for extended
version including heat supply and fuel input for 2003).
Electricity balance-sheet

Electricity consumption
Grid losses
Total required

Supply from electricity production companies


Supply from other electricity producers

2002

2003

Change

GWh

GWh

2002 - 2003

104,160
4,228
108,388

105,535
4,282
109,817

1.3%

62,866
29,140

63,973
28,849

-1.0%

1.8%

ow n decentralized production

12,046

12,261

1.8%

CHP and other supply to public grid

17,094

16,588

-3.1%

Import balance
Available domestically

of which
From public grid
From own decentralized production

From public grid


of which
Supply from electricity production companies

16,382
108,388

16,995
109,817

96,342
12,046
108,388

97,556
12,261
109,817

96,342

97,556

62,866

conventional supply
DH supply
CHP supply

Supply from other electricity producers

63,973
51,055

52,036

1.9%

7,871

7,753

-1.5%

4,184

5.8%
-7.6%

3,940

17,094

CHP supply

3.7%

16,588
11,955

11,107

WI supply

1,905

2,040

6.6%

"non fossil" supply

1,289

1,626

20.7%

1,815

-7.2%

gas engine supply

1,945

Import balance

16,382

From own decentralized production

12,046

16,995
12,261

production from ow n gas engines

2,981

3,095

3.7%

other generation

9,065

9,166

1.1%

Table 3.1: Electricity balance sheet for the Netherlands for 2002 and 2003 (Source: CBS, own
information)

35

Part 1

The electricity supplied by electricity production companies in Table 3.1 for 2002 and 2003 came from
the units also operating under the OVS agreement (central electricity production). This is evaluated
based on the production companies 2002 and 2003 Sustainability Reports for each location which
state the generated electricity and heat (including fuel amounts used and emissions produced). The
summarized generated electricity of these reports is slightly higher compared with the numbers in
Table 3.1 (1.4% for 2002 en 0.7% for 2003). It is checked with CBS (2006) and therefore clear which
units involve the CBS publication and this is one of the reasons why we use the CBS structure in the
further analysis and model calculations.
Table 3.1 shows growth in electricity consumption (GWh) but it is the electrical load that is relevant to
the balance between supply and demand. This is the instantaneous demand for electricity in the
Netherlands in MWe. One frequently displayed curve in the electrical load in the Netherlands is the
TenneT curve. The curve of the maximum load is published annually in TenneTs transport balance
sheet, but it shows only about 80% of the total Dutch load (situation as at 2003).
The TenneT curve is the sum of the electrical loads of all the observed (measured) production units
which supply electricity to the public grid4 plus any import/export at the time. This curve does not
incorporate the large contingent of industrial CHP units, small-scale CHP (gas turbines), waste
incinerators (WI), renewables etc. Figure 3.1 shows the TenneT curve for the maximum load on
Wednesday, December 12, 2001. Also included is an estimate of the load of the capacity not
observed by TenneT consisting of the individual consumption of sites and the capacity supplied to
the public grid. Figure 3.2 shows the same information for Saturday, January 26, 2002. Estimates are
necessary because no publications are available showing the total Dutch demand curve.
20,000

18,000

16,000

Load [MWe]

14,000

12,000

10,000

8,000

6,000

4,000
Observed by TenneT

2,000

Not observed
Total load

0
0

10

12

14

16

18

20

22

24

Time [hours]

Figure 3.1: Electrical load on a weekday (Source: TenneT, own information)


20,000

18,000

16,000

Load [MWe]

14,000

12,000

10,000

8,000

6,000

4,000
Observed by TenneT

2,000

Not observed
Total load
0
0

10

12

14

16

18

20

22

24

Time [hours]

Figure 3.2: Electrical load in a weekend (Source: TenneT, own information)


4

The net generation capacity of some large CHP units is measured from which their own on-site consumption is
still to be deducted. As this applies to both the demand and the supply curve it makes no difference when
assessing the load curve.

36

Chapter 3 Security of Supply

3.4

Electricity Supply

The total installed capacity in the Netherlands at the end of 2002 was approximately 20,100 MWe.
Compared with 2000 the installed capacity dropped by approximately 500 MWe due to the shutdown
of Lage Weide 5 and the mothballing of Flevo 30 in early 2001. NB: The gas turbine of the latter is be
re-used for peak load after installing an extra chimney behind the gas turbine.
The total installed capacity for 2003 increased to approximately 20,400 MWe mainly as a result of the
extra installed wind capacity in 2003. Combined with the import capacity of 3,350 MW (TenneT,
Transmission on Balance 2002 and 2003), this gives a capacity of approximately 23,500 MW in 2002
and approximately 23,700 MW in 2003. As stated before, around 80% of this capacity is observed by
TenneT and has been included in the TenneT load curve. Table 3.2 contains a breakdown of installed
generating capacity and import for both 2002 and 2003.

Generation + import
Generation capacity available
Conventional coal
Conventional gas
Hot windbox CCGT
Nuclear
Peak, Blackstart
IGCC
CCGT
CCGT district heating
Industrial CHP
Small-scale CHP
Wind turbines
Hydro electric plants
Solar PV systems
Biomass plants
Waste incineration plants
Others
Total generating capacity
Import
NEA - contracts (SEP)
Annual contracts
Monthly contracts
Daily contracts
Total import
Total generating capacity + import

2002

2003

MWe
3,960
2,100
2,308
449
168
253
1,826
2,000
4,273
1,539
670
37
26
27
414
67

%
16.9
8.9
9.8
1.9
0.7
1.1
7.8
8.5
18.2
6.6
2.9
0.2
0.1
0.1
1.8
0.3

MWe
3,960
2,100
2,308
449
168
253
1,826
2,000
4,316
1,513
906
37
46
27
414
67

%
16.7
8.8
9.7
1.9
0.7
1.1
7.7
8.4
18.2
6.4
3.8
0.2
0.2
0.1
1.7
0.3

20,117

85.7

20,390

85.9

900
900
849
701

3.8
3.8
3.6
3.0

900
900
849
701

3.8
3.8
3.6
3.0

3,350

14.3

3,350

14.1

23,467

100.0

23,740

100.0

Table 3.2: Total generation and import capacity in the Netherlands at the end of 2002 and 2003
(Source : CBS, TenneT, own information)
Figure 3.3 shows the increase in maximum load, as observed by SEP/TenneT, from 1995 till 2003.
The figure also shows the installed observed capacity (available) over 1995 - 2003. The sharp rise in
maximum load in 2001 compared with 2000 is explained by the increase in the generation capacity
observed in 2001.

37

Part 1

16,000
15486
15200

15046

15200

15200

15,000
14590

Maximum load, Installed capacity [MWe]

14447

14469
14238

14282

14242

14011
14,000

Max.load

13,000

Installed cap.

12558
12326
12055
12,000

11785
11360
11160

11,000

10,000
1995

1996

1997

1998

1999

2000

2001

2002

2003

Figure 3.3: Maximum load compared with installed observed generation capacity (Source: SEP/
TenneT, EnergieNed)
Despite the fact that more domestic units are included in the maximum load, the additional installed
capacity observed in 2001 compared with 2000 is only about 50% of the increase in the maximum
observed load. This result might be explained by the shutdown of a number of units (Lage Weide 5,
Flevo 30). Plainly, the growth in installed capacity till 2003 has not kept pace with the growth in
maximum load. The Dutch government decided that Borssele nuclear power plant (449 MW) can
continue operation until 2033. Building plans do exist for new units but with the exception of the
realization of Intergen and an industrial CHP at Shell Pernis (situation 2007) no steps have yet been
taken towards implementation5. There are also plans to put decommissioned capacity back into
operation if necessary. These pertain mainly to gas-fired conventional units with fuel efficiencies of
approximately 40%, which have been out of service for some years (ANP, 2005). We do not know
whether this is possible from a technical, organizational or legal perspective. Therefore after 2003 the
installed observed capacity has only improved modest.
3.5

Matching Supply and Demand

Figure 3.4 shows how the observed (available) generation capacity in the Netherlands relates to the
average E-day load and maximum load observed by TenneT. The decision to use the observed
generation capacity as a basis stems mainly from the approximately 80% share in the total generation.
The units not included in the observation play scarcely any part in matching supply and demand.
Because of the obligations to supply heat (steam), industrial CHP will not be shut down and will
continue to operate mostly at full capacity for the present as a result of the contribution relief and the
risks and costs of boiler operation. Small-scale CHP capacity, feeding to the public grid, mainly follows
a day/night schedule. The main task of waste incineration plants is to process as much waste as
possible; and maximum use is made of renewable sources to generate renewable energy.

Beside these new units new capacity is also added by the other electricity producers (i.e. wind, small scale CHP
etc). These units increase the total domestic installed capacity, however, it is unlikely that these units are part of
the installed observed capacity as shown in Figure 3.3 (see Section 9.3 of Chapter 9 for the latest developments
in new capacity for the electricity production companies and other electricity producers).

38

Chapter 3 Security of Supply

A number of important factors should therefore be taken into account when comparing the observed
generation capacity with the average and maximum load:
1) Average availability of units.
2) Mandatory deployment of units for, e.g., blast furnace gas, district heating, and the supply of
heat to industry. The deployment sequence is a static approach based on the average gas
and coal prices for 2003.
3) Capacity required for UCTE agreements and control and reserve capacity.
The capacity required by the market players for their own imbalance position is not included, nor are
the possible effects of grid restrictions or the differences between demand and supply in the summer
and winter periods.
A band of 85% - 95% is regarded as average availability of the observed generation capacity. An initial
estimate of the reduction of electricity generation due to heat supply is approximately 200 MWe. The
capacity required for UCTE agreements is 120 MWe for primary control and approximately 600 MWe
for secondary control and reserve capacity (this capacity cannot be used to supply the consumer
market).
Avr. Daily Load

20,000

Max. Daily Load


18,000

Gen 85% - H, RRV,UCTE


Gen 95% - H, RRV,UCTE

16,000

Gen 85% Available


Gen 95% Available
Gen
12,000

Gen + Import

10,000

Peak, Blackstart GT

18550

14469

HW CCGT
13000

6,000

12000

12920

14440

8,000

13520

Conv. Gas + Others


15200

Supply, Demand [MWe]

14,000

CCGT DH
CCGT
Conv. Coal

4,000

Import
2,000

WI
CCGT DH

D
LO
A
AV
R

LO
AX
M

85
%
G
EN

AD

O
RR
-C

O
RR
-C
95
%

G
EN

G
EN

85
%

AV
B

AV
B
95
%

G
EN

IM
P
+
G
EN

G
EN

DE
P

L.

SE
Q
U

EN
CE

Ind. CHP
Nuclear
IJmond, Velsen

Figure 3.4: Deployment sequence and generation capacity, comparison with average and maximum
load (Source: CBS, TenneT, own information)
At an average load of 13,000 MW (average estimate of maximum load over the whole year based on
annual demand) approximately 3,870 MW (13,520 + 3,350 (import) 13,000) of additional generation
capacity is still available at levels of 95%, while this works out at approximately 2,350 MW at levels of
85%. This additional capacity is a result of import capacity, which stood at 3,350 MWe at the end of
2003. Without import at 95% availability the additional capacity is just 520 MW; an 85% availability
results in a capacity shortfall of 1,000 MW.
To provide cover in periods bordering on maximum load (14,469 MWe in 2003) 2,400 MW of
additional generation capacity is available at levels of 95% while 880 MW is available at levels of 85%
(without import this results in shortfalls of 950 MW and 2,470 MW respectively). In the event of
emergencies and lower average availability of generating plants (85%) the outage of one large 600
MW unit can probably be absorbed by the primary and secondary control systems. In the case of

39

Part 1

major emergencies, such as the outage of two or more large units, not enough capacity is available.
This is because much of the import capacity has already been used up and it takes time to start up
units (start-up time for cold/hot start (in hours), peak load of gas turbines: 0.3/0.15, Hot windbox
CCGT: 5.5/2.5, CCGT: 2/1 and coal: 6/2.5). In such cases, security of supply is no longer guaranteed
and all the emergency and breaking capacity, if available, will have to be deployed to continue to meet
the demand. Given the strong links with the European grid, complete outage is technically unlikely, but
the imbalance prices may still soar. If this happens, some market players may be unable to honor their
commitments and (if these situations are chronic) certain customers might be disconnected.
At the end of 2003, the reserve factor, i.e. the generation capacity with or without import, divided by
the maximum load in the Netherlands, was 1.29 including import and 1.05 excluding import. The
reserve capacity therefore amounts to approximately 29% with import and approximately 5% without
import.
3.6

Summary

In this chapter the balance between electricity demand and generation capacity in the Netherlands is
analyzed for the reference situation in 2003. The installed observed capacity at average availability of
the units is compared with the average electricity demand and it is explained how this affects security
of supply. Finally reserve factors with and without import are calculated for 2003.

40

References

References
Algemene Rekenkamer (2004), Groene Stroom, Brief aan de tweede kamer der Staten-Generaal,
Vergaderjaar 2003-2004, 29630 no. 1-2 [Green Electricty, Letter to the House of Representatives
Parliamentary year 2003-2004]
ANP Pers Support, 2005, Essent wil elektriciteitcentrale Amer 7 weer in gebruik nemen, [Essent plans
to re-use the electricity plant Amer 7]
Beune, R.J.L. et al. (2004), Assessment of system adequacy: a new monitoring model, Cigre 2004.
CBS (2005), Duurzame energie in Nederland 2004 [Renewable energy in the Netherlands 2004]
ECN (2003), The Dutch Renewable Electricity Market in 2003,An overview and evelauation of current
changes in renewable electricity policy in the Netherlands, ECN-C--03-037
EnergieNed / SEP (1999), Elektriciteit in Nederland 1998 [Electricity in the Netherlands 1998]
EnergieNed (1999), Energie in Nederland 1999 [Energy in the Netherlands 1999]
EnergieNed (2000), Kerngegevens ten behoeve van de Energiesector 2000 [Key data in behalf of the
Energy sector 2000]
EnergieNed (2002), De energievoorziening in goede handen, Eerste bevindingen liberalisering
energiemarkt [The energy supply in good hands, first findings liberalised energy market]
European Commission (2002), 2002/358/CE Council Decision of 25 April 2002 concerning the
approval, on behalf of the European Community, of the Kyoto Protocol to the United Nations
Framework Convention on Climate Change and the joint fulfilment of commitments thereunder
European Commission (2001), Directive 2001/77/EC of the European Parliament and of the council of
27 September 2001 on the promotion of electricity produced from renewable energy sources in the
internal electricity market
European Commission (2000), Green paper Towards a European strategy for the security of energy
supply, European Commission, November 2000
Interview with CBS (2006)
Ministry of Housing, Spatial Planning and the Environment (2002), Evaluatienota Klimaatbeleid, De
voortgang van het Nederlandse klimaatbeleid: een evaluatie bij het ijkmoment 2002, [Evaluation
Report on Environmental Policy]
RIVM (2004), Milieubalans 2004 [Environmental balance 2004]
RIVM (2005), Netherlands National Inventory Report 2005 (NIR 2005), RIVM report 773201009/2005
SEP (1996), Elektriciteitsplan 1997 - 2006 [Electricity plan 1997 - 2006]
SEP (1996), Toelichting bij het Elektriciteitsplan 1997 - 2006 [Comments by the electricty plan 19972006]
TenneT (2000), Capaciteitsplan 2001 - 2007 [Capacity plan 2001 2007]
TenneT (2002), Capaciteitsplan 2003 - 2009 [Capacity plan 2003-2009]
TenneT (2002), Summarize out Ground rules concerning primary and secondary control of frequency
and active power within UCTPE) concerning the primary and secondary control

41

References

TenneT (2000), Transportbalans 2000 [Transmission on Balance 2000]


TenneT (2001), Transportbalans 2001 [Transmission on Balance 2001]
TenneT (2002), Transportbalans 2002 [Transmission on Balance 2002]
TenneT (2003), Transportbalans 2003 [Transmission on Balance 2003]
TenneT (2004), Transportbalans 2004 [Transmission on Balance 2004]
TenneT (2005), Transportbalans 2005 [Transmission on Balance 2005]
UCTE (2006), Physical electricity exchanges 2003*
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2002) Environmental behavior of the Dutch electricity
supply industy, Energie Techniek 7/8 July/August 2002
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2002) Price developments in a liberalized Dutch
electricity market, Energie Techniek 10 October 2002
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2002) Security of supply of electricity in the
Netherlands , Energie Techniek 9 September 2002
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2004) Trading between countries improving the balance
between Environment, Economy and Security of Supply or Power Play? The EU situation, Paper
presented at 19th World Energy Congress, Sydney, Australia, September 5-9, 2004
Van Eck (2007), A new balance for the energy sector. No longer a puppet in the hands of technology,
public interests and market, ISBN 978-90 788 8901-4
VROM (2002), Evaluatienota Klimaatbeleid, De voortgang van het Nederlandse klimaatbeleid: een
evaluatie bij het ijkmoment 2002 [Evaluation Report on Environmental Policy])

Internet Addresses
1.
2.
3.

42

http://www.milieubalans.nl
http://www.milieuennatuurcompendium.nl
http:/statline.cbs.nl

Part 2

Description of the System

44

Chapter 4 Energy Supply Chain and Supply Options

Chapter 4
4

Energy Supply Chain and Supply Options


4.1

Introduction

The balance between environment, economics and security of supply is determined by the complete
energy chain, including supply, transport/distribution, consumption by customers, and various external
factors that exert influence. Chapter 4 covers the actual energy supply chain or system, focusing on
the supply side of the service. The studied supply options are used as input for Part 4 and 5, the
scenarios and simulation model for the Dutch Electricity Supply. The studied supply options in this
chapter also serve as the basis for Part 3, Chapter 6: The technological developments of the supply
side. Chapter 5 describes energy transport and distribution to the users and discusses technological
developments in general terms.
4.2

Energy Supply Chain Overview

coal

export

import

Figure 4.1 shows the physical Dutch energy supply chain including the electricity generation options.

uranium
central electricity
generation

transmission
network

distribution
networks

other fuels
transport
network

consumption

gas
transport
network

distribution
network

other fuels

decentral electricity
generation

distribution
networks

renewable energy (wind, hydro, sun)

Figure 4.1: Graphic representation of the physical Dutch energy supply chain
A large diversity can be seen in the electricity generation options due, amongst others, to:
Differences in primary energy sources relating to application, costs, applicable regulations
Differences in technologies
Presence of industrial heat users and/or district heating users
Regulations, laws.
4.3

Definition Supply Side

4.3.1

Main Categories

The diversity in electricity generation in the Netherlands for this research is divided into three main
categories (see also Figure 4.2) based on the differences in energy conversion technologies which
use fossil fuels, the use of high, medium, and low temperature heat for heating purposes in the
industry and households and renewables:

45

Part 2

Conventional and combined cycle generation


Combined heat and power (CHP) or Cogeneration:
o Industrial Heat/Power
o District Heating
Renewable generation.

RENEWABLES (1,016)

IND CHP (4,316)

CONV (11,545)
DH CHP (3,513)

Figure 4.2: Installed capacity (in MWe) in the Netherlands by category by the end of 2003 (Source:
CBS, Cogen)
Together, these supply options represent a total of 20,390 MWe in installed capacity in the
Netherlands (end of 2003; see also Part 1, Chapter 3), of which approx. 57% conventional capacity,
approx. 38% CHP (industrial, heat distribution), and 5% non-fossil.
The three main categories conventional and combined cycle generation, CHP, renewable generation
of the supply options will be studied in more detail in Section 4.4, 4.5, and 4.6. A split-up based on
applied technology of the main categories is provided in Figure 4.3.

BIOMASS (27)
HYDRO (37)
WIND (906)

SOLAR PV (46)
OTHERS (67)

WASTE INC (414)

CONV COAL (3,960)

SMALL CHP (1,513)

CONV GAS (2,100)

IND CHP (4,316)


HW CCGT (2,308)

CCGT DH (2,000)

CCGT (1,826)

IGCC (253)

NUCLEAR (449)

PEAK / BLACKSTART (168)

Figure 4.3: Supply options (in MWe) in respect of electricity generation in the Netherlands by the end
of 2003

46

Chapter 4 Energy Supply Chain and Supply Options

The largest share, in terms of capacity, is for combined heat and power units (industrial and district
heating), followed by conventional coal-fired units, hot wind box CCGT units, conventional gas-fired
units, and CCGT units without heat supply.
4.3.2

Basic and Combined Cycles in Energy Conversion Technology

Energy conversion technology is based on different conversion processes. Chemical energy is


converted into heat which is transferred to a working fluid. This fluid undergoes a thermodynamic cycle
in which thermal energy is converted into mechanical energy. For more insight on the main future
developments in electricity generation technology (Chapter 6) first a brief look at the most applied
thermodynamic and combined cycles and their main characteristics is given.
Basic Thermodynamic Cycles
Thermodynamic cycles in the generation of electricity are: Rankine, Brayton/Joule, Otto, Diesel,
Stirling, and Ericsson. The Rankine, Brayton/Joule and Otto/Diesel cycles are used on a large scale.
Their main differences are in the temperatures and pressures on which they operate en less by the
mechanical design and by the limits of their applied materials. These differences will be discussed in
more detail in Section 4.4 and 4.5. For a detailed explanation of the other cycles see Korobitsyn
(1998) and other relevant literature.
Combined Cycles
Basic thermodynamic cycles have intrinsic limitations and cannot operate within a broad temperature
range. Therefore no single cycle can offer high efficiencies but combining single cycles with a hightemperature topping cycle and a medium- or low-temperature bottoming cycle can obtain a better
performance. In Figure 4.4 topping and bottoming cycles are ranked according to their temperature
range in order to determine what cycle is better suited for topping or bottoming application.

Figure 4.4: Thermodynamic cycles arranged according to their temperature range (Source:
Korobitsyn, 1998).
Combining high-temperature cycles with medium- and low-temperature ones provides the most
effective way in approaching Carnot efficiency, and thus better utilization of fuel exergy.
A number of combinations of cycles can be distinguished (see Table 4.1).

47

Part 2

Topping cycles

Bottoming cycles

Rankine

Otto/Diesel

Joule

Fuel cell

Rankine
Kalina
Joule
Otto/Diesel
Stirling
Fuel cell
Heat pump

Table 4.1: Combinations of cycles (Source: Korobitsyn, 1998)


The Brayton/Joule (gas turbine) cycle, the Otto/Diesel cycle, and the high-temperature fuel cells can
be applied as topping cycles because of their relatively high exhaust temperatures. The Rankine and
Stirling cycles are suitable both for topping and bottoming. However, various factors such as the status
of development, power output, fuel requirements, or part load characteristics can limit the possibilities
of combination. According to Korobitsyn (1998) the Joule/Rankine cycle, Joule/Joule cycle and the
Joule cycle with a topping high-temperature fuel cell are of interest for future development.
Future developments for this research
Now recently the Fuel Cell and Stirling cycle get more attention because of their applicability on a
small scale or by expected higher conversion efficiency. Research in the exotic cycles continues but
this is beyond the scope of this research as it is assumed that large scale implementation is not likely
within the next 30 years. Here we concentrate on the development of existing large scale applications
like the Rankine and Joule/Rankine cycles and those applications that have a potential for large scale
applications within the next 30 years (see Chapter 6) as they are crucial for the electricity supply
model (see Part 4).
4.3.3

Definition of Input and Output Parameters for the Model

For the various supply options displayed in Figure 4.3 an overview will be provided of installed units by
category (conventional and combined cycle, large (> 100 MWe) CHP, small CHP and renewables as a
cluster. Only in the case of large-scale units, a description of the technology is given (applied cycles
and main characteristics, applied fuels, production process, flue gas cleaning, and residual products (if
applicable)). A summary of all main characteristics of the supply options for use in the electricity
supply model is given in Section 4.7; the following data is presented graphically or in table form, by
supply option:
Technology: heat consumption curve, start-up times, minimum up- and downtime,
controllability, maintenance philosophy
The environment: required/possible environmental measures, emissions and residual
products, process integration options (for example, indirect co-firing of biomass, steam from
another unit, gasification), need for cooling water, site requirements, environmental costs
Security of supply: permits and permit terms, time to build including necessary preparations,
investment and operating costs, dismantling costs.
These data will be used as input and output parameters for the Dutch electricity supply model and as
a basis for defining the scenarios (see Part 4).

48

Chapter 4 Energy Supply Chain and Supply Options

4.4

Conventional and Combined Cycle Units

Generally, every conventional unit, based on the classical Rankine cycle, comprises a heat source,
boiler, working fluid, steam turbine and condenser. Every combined cycle unit comprises a gas turbine
as heat source, waste heat boiler, working fluid, steam turbine and condenser. Both cycles also use
derived components such as cooling water facilities and a cooling water source (sea / river water,
cooling towers), and a step-up transformer to transform the electricity generated by the generator(s)
(usually 10kV) to the required voltage of the high-voltage connection which is used by the unit. The
conventional and combined cycle units are described according to their share of total generation.
4.4.1

Conventional Coal-Fired Units

Installations
Present-day (2008), seven conventional coal-fired units at five locations are operational in the
Netherlands (see Table 4.2), of which three (G13, A-81, A-91) deliver heat.

Name or location

Amer 8
Gelderland 13
Borssele 12
Maasvlakte 1
Maasvlakte 2
Amer 9
Hemweg 8

Label

A-81
G-13
BS12
MV-1
MV-2
A-91
HW-8

Owner (present)

Essent
Electrabel
EPZ
E.ON
E.ON
Essent
Nuon

Commisioning

1980
1981
1987
1988
1988
1993
1994

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

Temperature
[C]

primary

645
602
403
520
520
600
630

39
39
39
39
39
41
42

182
185 / 38
183
183
183
270 / 58,5
260 / 50

543 /
540 /
543 /
540 /
540 /
540 /
540 /

Coal
Coal
Coal
Coal
Coal
Coal
Coal

543
540
543
540
540
568
568

------ Fuel type -----secondary


tertiary
Gas / Oil
Oil
Gas
Gas
Gas
Gas
Gas

Biomass
Waste Wood
Biomass
Biomass
Biomass
Biomass (gas)

Table 4.2: Coal-fired units (Source: SEP, own information)


Technology
Applied Cycle
All conventional coal fired units use an advanced Rankine Cycle with superheat and singular reheat of
which pressures and temperatures are given in Table 4.2. The applied temperatures for superheat do
not differ a lot between the units in contrast with the applied pressures. In the two latest built units,
Hemweg 8 and Amer 9, a super-critical steam boiler is applied with a boiler outlet pressure of 260 270 bar.
Fuel Supply
All units use coal as primary fuel but for start-up they use natural gas or oil. Direct co-firing of biomass
or biogas is also possible.
Coal
The coal intended for coal-fired units is delivered by ship from all over the world, sometimes via a
transshipment agent. Various types of coal with their one composition are collected, mixed and sent
by bulk conveyor to a circular storage facility. Here the coal is picked up by a scraper and transported
to the power station via bulk conveyors and weighing machines. All the conveyor belts are enclosed to
prevent dust and noise and are fitted with sophisticated fire-alarm and fire-extinguishing systems.
Precautionary measures are also taken to prevent soil pollution; the circular storage facility has a
watertight foundation to prevent rain and contaminated spray from penetrating the soil.
Natural Gas
The natural gas for units that use gas as the secondary fuel is transported under high pressure via the
Gasunie Transportservices (GTS) pipelines to custody transfer stations at the sites in which the gas
pressure can be adjusted before the gas is transported to the boiler burners.
Oil
Oil is transported by ship and stored on site in tanks. In the event of a coal shortage, for example due
to extreme winter conditions that disable transport by ship, these tanks usually have enough oil

49

Part 2

storage capacity to keep the power station operational for two weeks (in the case of G13). Oil pumps
are used to transport the oil from the tanks to the boiler burners. Heat tracing is applied to prevent the
oil to stall in the pipelines during freezing conditions.
Biomass
As a result of agreements between the government and the electricity producers to reduce emissions
of CO2 and other greenhouse gases, the sector is aiming to collectively reduce CO2 emissions by 5.8
Mtons a year by 2012 for the same electricity production. This target is to be realized by substituting
part of the coal requirement with biomass or organic waste. In Gelderland 13 this is achieved by
grinding chips from purified waste wood to a powder, which is blown into the boiler. At Amer 9 it takes
place via an interim step involving wood gasification. The Maasvlakte units burns besides biomass
also organic waste. The other units have plans for co-firing of biomass.
Production Process
In conventional power plants the fuel is fed straight into the steam boiler. In conventional coal-fired
plants the primary fuel is obviously coal, but some are also able to run on natural gas, oil, or on
biomass which amounts approximately 20% for biomass with a percentage of moisture of 15%
(Novem, 2001). Figure 4.5 shows the production process at a coal-fired power plant.

Figure 4.5: Diagram of the production process at a coal-fired plant (Source: PGEM, 1970s)

50

Chapter 4 Energy Supply Chain and Supply Options

The coal is ground (pulverized) to a fine powder in huge grinders, so that less than 2% is +300 m and
70-75% is below 75 m, for a bituminous coal. The pulverized coal is blown with part of the
combustion air (primary air) into the boiler through a series of burner nozzles. Besides the primary air
also secondary and tertiary air are added as combustion air. Combustion takes place at temperatures
from 1300 - 1700C, depending largely on coal rank. Particle residence time in the boiler is typically 2 5 seconds, and the particles must be small enough for complete combustion to have taken place
during this time. Boiler feed water is pumped first into the economizer and after being preheated it is
led through the evaporating sections in the furnace walls of the boiler. Due to the radiation heat of the
flames the boiler feed water is converted into steam and after the steam/water separators the steam is
overheated in the superheat sections due to convection heat. Steam pipes lead the steam to the high
pressure part of the steam turbine in which the steam expands to middle pressure and is reheated
again in the boiler and again lead via steam pipes to the middle pressure part of the steam turbine.
The steam then expands in the middle pressure part of the steam turbine to low pressure conditions
and via the low pressure part of the steam turbine to very low pressure corresponding to the vapor
pressure of the temperature in the condenser (almost vacuum conditions). Meanwhile, the energy in
the steam is converted first into mechanical energy and then via a generator into electrical energy.
The expanded and exhausted steam is cooled in a condenser (a large vessel with cooling water
flowing through thousands of pipes) and converted back to water. This condensate is then pumped by
a condensate pump to the accumulator or deaerator. De-aereation removes, amongst others, traces of
oxygen from the feed water to prevent corrosion in the steam turbine etc. After passing through the
deaerator the feed water is brought up to pressure with feed pumps and returned to the boiler where
the process starts all over again.
The boilers at Hemweg 8 and Amer 9 are designed for super-critical pressures. At full capacity the
pressure on the outlet of the Hemweg boiler is 260 bar. Under super-critical pressure the water is
immediately converted into steam; so there is no vaporization phase. Compared with the other non
super critical units the electrical efficiency is around 3 - 4% higher. The condenser pressure has also
influence on the electrical efficiency. A lower cooling water temperature means a lower condenser
pressure and therefore more expansion of steam and more enthalpy drop (and power output) in the
lower pressure part of the steam turbine. Units close to the sea normally have an advantage on units
at a river of units which have natural draft cooling towers.
Flue-Gas Cleaning and Residues
Coal combustion leads to emissions of, amongst others, NOx, fly ash and SO2, which, if left untreated,
are harmful to the environment. Hence, the flue gases from coal-fired units are cleaned as much as
possible. This takes place as follows:

The use of Low-NOx burners in conjunction with a furnace of optimal dimensions.These burners
bring about low NOx concentration in the flue gases. Denitrification (DeNOx): this is based on a
chemical process in which ammonia is added to the flue gases so that the nitrogen oxide is
converted into water vapor and nitrogen. A catalyst is needed to trigger this chemical process.

Electrostatic fly ash filters: fly ash particles (approximately 90 - 95% of the produced ash)
released by the burning of pulverized coal are caught in electrostatic filters and are then used as
a raw material in the building and cement industry.

Flue gas desulphurization: around 90% of the sulphur-dioxide is removed from the flue gases by
applying the wet limestone process in which a suspension of finely produced limestone or lime
is sprayed countercurrent to the flue gases. The SO2 from the flue gas reacts with the limestone
to form a gypsum suspension which is further thickened and purified by hydrocyclones and
centrifuges. The end product is high-grade powdered gypsum.
Other Residues
Around 5 - 10% of the produced ash is bottem ash. This is used as an additive in the manufacture of
cement and in the road-building industry.
Heat Supply
Dutch coal-fired units are all situated in densely-populated residential or industrial areas. The two coal-fired
units of the Amer centrale already supply greenhouse and district heat on a large scale, Nijmegen supplies
process heat (steam) to a limited degree, and in Amsterdam the Hemweg 8 is located close to the
infrastructure of a district heat project.

51

Part 2

4.4.2

Conventional Gas-Fired Units

Present-day, four conventional gas-fired units distributed over two sites are operational in the
Netherlands (see Table 4.3).

Name or location

Velsen 24
Claus A
Claus B
Velsen 25

Label

VN24
CC-A
CC-B
VN25

Owner (present)

Nuon
Essent
Essent
Nuon

Commisioning

1974
1977
1977
1986

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

Temperature
[C]

primary

459
640
640
361

40
41
41
43

180 / 40
257,5 / 48,5
257,5 / 48,5
180 / 40

545 /
540 /
540 /
530 /

BFgas
Gas
Gas
BFgas

540
540
540
530

------ Fuel type -----secondary


tertiary
Gas
Oil
Oil
Gas

Oil
Bio-oil
Bio-oil
Oil

Table 4.3: Conventional gas-fired units (Source: SEP, own information)


The two Velsen units differ from the Claus units in that they burn residual gases like blast furnace gas,
cokes oven gas, and oxy gas produced by the manufacturing of steal by Corus (previously
Hoogovens).
Technology
Applied Cycle
All conventional gas fired units use an advanced Rankine Cycle with superheat and singular reheat of
which pressures and temperatures are given in Table 4.3. In Claus A & B super critical steam boilers
are applied with a boiler outlet pressure of 257.5 bar.
Fuel Supply
Blast Furnace- , Cokes oven and Oxygas
Residual gases which are released in the steel production process comprise blast furnace gas, cokes
oven gas, and oxy gas. The boilers at Velsen 24 and 25 (Corus) are equipped to stoke a maximum
capacity of mixed blast furnace gas. Natural gas is supplementary used and in can also be used in
case no residual gases are available.
Natural Gas
The natural gas is transported under high pressure via the Gasunie Transportservices (GTS) pipelines
to custody transfer stations at the sites in which the gas pressure can be adjusted before the gas is
transported to the boiler burners.
Oil
See section on conventional coal.
Bio oil
At the Claus power station bio-oil is used in addition to natural gas (the main fuel). The treatment is
similar to the treatment of oil as described in the section on conventional coal.
Production Process
The production process for conventional gas-fired units is similar to the process for coal-fired units.
The only thing that is missing is the coal treatment and the flue-gas purification. The natural gas or the
combination of natural gas and blast furnace gas is combusted in a steam boiler with the addition of
combustion air.

52

Chapter 4 Energy Supply Chain and Supply Options

4.4.3

Hot Windox Combined Cycle Gas Turbine (HW CCGT)

In the Netherlands there are present-day five so called hot windbox combined cycle gas turbine units
distributed over four sites. Four of the units originally were build as conventional gas-fired units on
which, during retrofit, a gas turbine was placed upstream (see Table 4.4). The Harculo 6 unit was
directly built as a HW CCGT. Two HW CCGT units, not mentioned in the table, are no longer
operational: Flevo 301 (conserved) and LageWeide 5 (dismantled).

Name or location

Eems 20
Bergum 10
Bergum 20
Harculo 6
Hemweg 7

Label

EC20
BG10
BG20
HC60
HW-7

Owner (present)

Electrabel
Electrabel
Electrabel
Electrabel
Nuon

Commisioning
[com / retrofit]

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

Temperature
[C]

primary

1977 /
1974 /
1975 /
1982
1979 /

695
332
332
350
599

45
45
45
44
44

180 /
180 /
180 /
170 /
180 /

540 /
540 /
540 /
540 /
540 /

Gas
Gas
Gas
Gas
Gas

1987
1986
1986
1988

40
40
40
40
40

540
540
540
540
540

------ Fuel type -----secondary


tertiary

Oil (only Boiler)


Oil

Table 4.4: Hot Winbox Combined Cycle units (Source: SEP, own information)
Technology
Applied Cycle
As stated above, four of the five HW CCGT units where conventional gas fired units use an advanced
Rankine Cycle with superheat and reheat and use a one-pressure steam boiler. During the years 1986
1988 a gas turbine was placed upstream and the cycle became a Joule/Rankine cycle where the
gas turbine exhaust is fed into the existing steam boilers furnace, this is also called repowering
scheme (Korobitsyn, 1998; Kehlhofer et al, 1999).
Fuel Supply
Natural Gas
The natural gas for units that use gas as the secondary fuel is transported under high pressure via the
Gasunie Transportservices (GTS) pipelines to custody transfer stations at the sites in which the gas
pressure can be adjusted before the gas is transported to the boiler and gas turbine burners.
Oil
See section on conventional coal.
Production Process
As far as boiler and steam turbine are concerned the production process for conventional gas-fired
units with upstream placement of gas turbines (for gas turbines see Section 4.4.7, CCGT) is similar to
the process for conventional gas-fired units (see Figure 4.6). However, in all the power stations
(except Harculo 60, which already included upstream placement) upstream placement of gas turbines
took place later, see Table 4.4.

The gas turbine at Flevo 30 has (2005) been converted into a stand-alone peak-load installation.

53

Part 2

Figure 4.6: Flow diagram of a typical HW-CCGT-unit (Source: Castanier et al, 1988)
The technology for the upstream placement of gas turbines stems from the need to make productive
use of the hot exhaust gases from gas turbines as pre-heated combustion air for the downstream
placed conventional steam boiler. The exhaust gases from the gas turbine still have an oxygen
content of 15% and this is still enough to serve as pre-heated combustion air for the burners of the
conventional steam boiler. With the combination of gas turbine and steam cycle the total energy output
of the installation is increasing by around 15%. Another advantage of generating electricity in HW
CCGT-units is the lower emissions of nitrogen oxide. This is, of course, partly due to the increase in
the electricity output (less gas required for the same electricity production) but also and more
importantly to the special Dry Low NOx combustion technology in gas turbines (see Section 4.4.7)
which emits less NOx and eases the load on the furnace and less NOx forming in the conventional
steam boiler because the combustion air has a lower oxygen content. All the installations can operate
in stand-alone steam boiler mode, stand-alone gas turbine mode, or a combination of the two. The
latter mode is the most common as it delivers the highest electricity output at a maximum electrical
efficiency. The gas turbines have their own generator and are separate from the conventional steam
boiler, and the hot exhaust gases are transported through a labyrinth of flue ducts to eventually arrive
at the burners of the conventional boiler.
4.4.4

Nuclear

Present-day there is one nuclear power plant in the Netherlands situated in Borssele (see Table 4.5).

Name or location

Borssele 30

Label

BS30

Owner (present)

EPZ

Commisioning

1973

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

Temperature
[C]

primary

449

32

150

300

Uranium

Table 4.5: Nuclear unit (Source: SEP, EPZ, own information)

54

------ Fuel type -----secondary


tertiary

Chapter 4 Energy Supply Chain and Supply Options

Technology
Applied Cycle
The nuclear unit uses a Rankine Cycle with a one-pressure (secondary) steam circuit.
Fuel
Uranium
Nuclear power plants use uranium as fuel. Natural uranium consists of 99.3% non-fissionable material
(U238) and only 0.7% fissionable material (U235). A nuclear power plant needs material with a higher
content of fissionable uranium U235 (e.g. 4% in Borssele). So, the uranium first needs to be enriched,
i.e. the percentage of fissionable U235 needs to be artificially increased. This takes place in gas
diffusion plants or ultra-centrifuges at, amongst other places, the Urenco enrichment company in
Almelo. Pellets the size of a fingertip are pressed from the enriched uranium oxide (UO2). Two
hundred and forty of these pellets are placed in a corrosion-resistant rod; two hundred and five of
these rods are assembled to form a nuclear fuel element. There are 121 of these elements in the
reactor vessel. This is where the nuclear fission takes place. Once a year the unit is shut down for a
few weeks so that the fuel can be changed. Around a quarter of the 121 elements are then replaced
by new ones. Thus, the entire load is renewed every four years. Borssele uses 10 tons of nuclear fuel
a year (the coal-fired 406 MWe unit Borssele12 used 981,443 tons of coal in 2003 (Source: EPZ,
2004)). The period in which the fuel is exchanged is also reserved for maintenance. The spent
elements are stored temporarily in a basin where they cool off before being loaded in containers and
transported to a reprocessing plant in France. Here the usable substances are extracted and the
remainder is vitrified and packed in corrosion-resistant cylinders. These cylinders eventually return to
the Netherlands for permanent storage.
Production Process
The nuclear power plant at Borssele (see Figure 4.7 and 4.8) showing the generator room (right) and
the domed reactor building (left).

Figure 4.7: Nuclear power plant with the generator room on the right and the reactor building on the
left (Source: Stichting Borssele 2004+)

55

Part 2

The generator room also called the conventional part houses the high pressure (6) and the low
pressure (7) steam turbines, the generator (8) with the exciter (9), the condenser and the accumulator
/ dearator tank (10) (see Figure 4.8). The dome accommodates the reactor vessel (1), the two steam
generators (2), the main coolant pump (3), the pressuriser (4) and the blowdown tank (5).

Figure 4.8: Diagram of the nuclear energy plant (Source: Stichting Borssele 2004+)
The reactor vessel is made of high-grade steel. It is approximately 10 meters high and 4 meters in
diameter with a wall thickness of 18.5 cm. The reactor vessel holds the 121 uranium fuel elements
enclosed by water.
The fission process involves hitting the nucleus with a neutron at a limited speed, using water as a
moderator (first function of the water): fast-moving neutrons collide with the hydrogen nuclei and lose
their speed. The U235 uranium atoms in the fuel are fissioned by the absorption of a neutron. This
gives rise to 2 fission products and 2 or 3 new neutrons. The fission products try to bolt away at very
high speed but are restrained because they are incarcerated in the pellets, producing large amounts of
heat at the same time. This heats up the rods, which in turn heat up the water (second function of the
water) in the reactor vessel. The water will either boil in a boiling water reactor (BWR) to generate
steam which is used directly in a steam turbine, or it will transfer the heat to a secondary (steam)
circuit (operating at lower pressure) to produce steam in a pressurized water reactor (PWR). The
newly formed neutrons can again split uranium atoms thus sustaining the chain reaction. A small
fraction of the neutrons can be absorbed by U238 uranium atoms which are then transformed to
Plutonium Pu239 which can also split.
Borssele is a PWR, where the water in the vessel is under high pressure (150 bar, 300C). The
thermal energy of the water in the reactor vessel is transferred by means of a heat exchanger to a
secondary circuit in which steam is produced. The turbine is driven by this produced steam.
The heat production in the reactor vessel is regulated using water as a moderator and with elements
containing a neutron-absorbing substance which can be inserted between the fuel rods. The elements
can be moved in and out to catch the desired amount of neutrons. The reaction process can thus be
slowed down or speeded up to reduce or increase the electricity output. If the elements are placed at
maximum depth between the fuel rods they will catch so many neutrons that the process will come to
a standstill.

56

Chapter 4 Energy Supply Chain and Supply Options

Waste
The spent uranium fuel used during a year of electricity generation in Borssele consist of unused
uranium 235 ( 1%), activated products (Plutonium (useful again) and other trans uranium elements)
and fission products. Around 95% of the spent uranium fuel can be re-used; so, 5% constitutes real
waste. This waste is radioactive, sometimes for a very long time (thousands of years). On average,
after reprocessing, 1.3 m3 of fission waste is left, 2-3 m3 of high-radioactive waste (including trans
uranium elements) and 30-50 m3 of low-mid radioactive waste. The latter category is no longer
dangerous after a few decades (EPZ, 2000). This waste is stored in various ways; it may, for example,
be placed deep under the ground in salt mines, or in layers of clay or in rock structures (Stichting
Borssele 2004+). The Dutch have opted for a temporary solution, namely, the surface storage facility
at COVRA (Central Organization for Radioactive Waste Borssele). Radioactive waste from other users
(e.g. hospitals) is also stored there.
4.4.5

Peak-Load Units

There are 8 peak-load units with gas turbines in the Netherlands at the moment spread over 4 sites
(see Table 4.6). One peak load unit, not mentioned in the table, is no longer operational: FLG1
(conserved). The units are also deployed as emergency or black-start capacity.

Name or location

Delft
Delft
Delft
Delft
Borssele
Amer
Amer
Velsen

Label

DE-1
DE-2
DE-3
DE-4
BS20
A-20
A-21
VNG1

Owner (present)

E.ON
E.ON
E.ON
E.ON
EPZ
Essent
Essent
Nuon

Commisioning

1974
1974
1974
1974
1972
1972
1972
1975

Net Power
[MWe]

Net E-efficiency
[%]

primary

23
23
23
24
18
15
15
26

26
26
25
26
25
26
26
26

Gas
Gas
Gas
Gas
Oil
Gas
Gas
Gas

------ Fuel type -----secondary


tertiary
Oil
Oil
Oil
Oil
Gas

Oil

Table 4.6: Peak-load units (Source: SEP, own information)


Technology
Applied Cycle
The peak load units use a Brayton/Joule cycle.
Fuel
Natural Gas
Except for the four gas turbines in Delft, each unit shares a site with other units using the existing gas
infrastructure. The Delft site is connected with GTS custody transfer station in which the gas pressure
can be adjusted before the gas is transported to the gas turbines.
Oil
Oil (mostly light oil) is transported by ship or truck and stored on site in tanks. Oil pumps are used to
transport the oil from the tanks to the gas turbine burners.
Production Process
The gas turbines are stand-alone installations, whereby the hot exhaust gases of 500C are led
directly through a chimney into the environment.

57

Part 2

4.4.6

Integrated Gasification Combined Cycle (IGCC)

At present the Willem Alexander Power station in Buggenum is the only IGCC unit in the Netherlands
(see Table 4.7). This IGCC came at the beginning of 1994 into operation as a demonstration project
and was the first European and at that time the world largest integrated coal gasification combined
cycle. Present the Willem Alexander Power station is a commercial operating power plant.

Name or location

Demkolec

Label

MC-7

Owner (present)

Nuon

Commisioning

1993

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

Temperature
[C]

253

43

125 / 29 / 9

510 / 240 / 185 Coal

primary

------ Fuel type -----secondary


tertiary
Gas

Biomass

Table 4.7: Integrated Gasification Combined Cycle unit (Source: SEP, Energietechniek, 1996, own
information)
Technology
Applied Cycle
The applied cycle for the Willem Alexander Power station are the Brayton/Joule cycle as topping cycle
and the Rankine cycle as bottoming cycle. The Rankine cycle is implemented as an unfired heat
recovery steam generator (HRSG) behind a Brayton/Joule cycle, the gas turbine.
Fuel Supply
The unit can operate both on syngas and on natural gas. The coal can also, partially, be replaced by
biomass which makes the unit a multi fuel unit.
Syngas
The coal supply, treatment and pulverization to a fine powder is the same as in conventional coal-fired
unit but from there on, the process is different and works on the basis of synthesized gas from a coal
gasifier. The gasification procedure was developed by Shell on the basis of oil gasification. The
process is shown in Figure 4.9.

Figure 4.9: Coal gasification process (Source: World Coal Institute, 2003)

58

Chapter 4 Energy Supply Chain and Supply Options

The principle is based on the entrained bed model with the addition of pure oxygen. In the entrained
bed gasification system both the bed material and the uncombusted particles are dragged upwards
out of the gasifier, so that the contact time between the fuel particles and the bed material is extremely
short, 2 - 4 s. Gasification takes place at a temperature of 1000 - 1500C and a pressure of 28 bar in
the presence of a controlled shortage of air/oxygen, thus maintaining reducing conditions. The
process is carried out in an enclosed pressurized reactor, and the product (called synthesis gas,
syngas or fuel gas) is a mixture of carbon monoxide (CO) and hydrogen (H2). The non combustible
fraction of the coal (minerals) coagulates on the reactor walls and comes out as slag. In this phase the
syngas still has several contaminations which are removed step by step. After gasification the syngas
is cooled with quench gas to lower the temperature to 900C. After cooling the syngas in a convection
cooler, in which the released heat is transferred into steam, the fly ash particles are removed. Next the
syngas is washed with water to remove the soluble fractions as chlorides and fluorides and the syngas
is desulphurized. The cleaned syngas is then led to burners of the modified KWU V94-2 gas turbine of
which the exhaust gases are led through a HRSG to produce steam. This steam is used along with the
steam from the convection cooler to drive a steam turbine. The oxygen and nitrogen needed for the
gasification process are produced in an air separation plant which is connected to the compressor of
the gas turbine from which part of the air flow is diverted.
Natural Gas
The natural gas for the unit is transported under high pressure via the Gasunie Transportservices
(GTS) pipelines to the custody transfer station at the site in which the gas pressure can be adjusted
before the gas is transported to the gas turbine burners.
Biomass
The coal can be replaced partially by biomass which undergoes the same gasification process. Up to
50% of biomass can be mixed with coal (Hotchkiss, 2002). Not all the coal can be replaced due to the
fact that a protective slag layer on the reactor walls during gasification has to be maintained and this
layer is formed mainly by the gasification of coal. Another reason is the lower calorific value of
biomass compared with coal. For the same electricity production more biomass is required and this is
limited due to maximum capacities of the existing coal treatment facilities of the unit.
Production Process
The applied cycle for IGCC are the Brayton/Joule cycle as topping cycle and the Rankine cycle as
bottoming cycle. The Rankine cycle is implemented as an unfired heat recovery steam generator
(HRSG) behind a Brayton/Joule cycle, the gas turbine. The CCGT part of the IGCC installation is
described in detail in Section 4.4.7.
Flue-Gas Cleaning and Residues
Pollutants from IGCC are removed from IGCCs high-pressure fuel gas stream (prior to combustion)
see Figure 4.9, rather than from the exhaust gas generated by total combustion in a PCC plant (postcombustion). Hydrogen sulfide (H2S) and carbonyl sulfide (COS), once hydrolyzed, are removed by
dissolution in an organic solvent and converted to valuable by products such as elemental sulfur or
sulfuric acid. The to ammonia (NH3) converted nitrogen, as well as some hydrogen cyanide (HCN) and
thiocynate is removed via water scrubbing. Most trace pollutants are removed with the slag/bottom
ash or in the particulate control equipment (Ratafia-Brown, 2002). Particular matter sulfur, nitrogen
and mercury are more effectively removed in IGCC than in conventional coal combustion systems like
PCC (see Figure 4.10). The emissions of particulates, NOx and SO2 from IGCC units are expected to
meet, and possibly to exceed, all current standards.

59

Part 2

Figure 4.10: Emissions differences between IGCC and PCC (Source: Nuon, 2006)

4.4.7

Combined Cycle Gas Turbine (CCGT)

There are at present 6 CCGT units in the Netherlands, spread over 2 sites (see Table 4.8).

Name or location

Eems
Eems
Eems
Eems
Eems
Donge

Label

EC3
EC4
EC5
EC6
EC7
DGS1

Owner (present)

Electrabel
Electrabel
Electrabel
Electrabel
Electrabel
Essent

Commisioning

1996
1996
1996
1996
1996
1976

Net Power
[MWe]

Net E-efficiency
[%]

Pressure
[bar(a)]

341
341
341
341
341
121

53
53
53
53
53
43

112.2 /
112.2 /
112.2 /
112.2 /
112.2 /
-

28.5 /
28.5 /
28.5 /
28.5 /
28.5 /

4.9
4.9
4.9
4.9
4.9

Temperature
[C]

primary

541 /
541 /
541 /
541 /
541 /
-

Gas
Gas
Gas
Gas
Gas
Gas

541 /
541 /
541 /
541 /
541 /

263
263
263
263
263

------ Fuel type -----secondary


tertiary

Table 4.8: Combined Cycle Gas Turbine units (Source: SEP, Energietechniek, own information)
Technology
Applied Cycle
The applied cycles are the Brayton/Joule cycle as topping cycle and the Rankine cycle as bottoming
cycle. The Rankine cycle is implemented as an unfired heat recovery steam generator (HRSG) behind
a Joule cycle, the gas turbine.
Fuel Supply
Natural Gas
The natural gas for the unit is transported under high pressure via the Gasunie Transportservices
(GTS) pipelines to the custody transfer station at the site in which the gas pressure can be adjusted
before the gas is transported to the gas turbine burners.

60

Chapter 4 Energy Supply Chain and Supply Options

Production Process
The CCGT installations (solely for electricity generation) consist of an industrial gas turbine(s), heat
recovery steam generator(s) and a condensation steam turbine(s) without steam extraction which are
described in more detail below.
Gas Turbine
An industrial gas turbine consists of a compressor (1), a combustion chamber (2), an expansion
turbine (3) and a generator, possibly connected via a gearbox, see Figure 4.11.

Figure 4.11: Single-shaft GE Frame 6 gas turbine (Source: Thomassen International)


Air is drawn in from outside via the compressor and passes through filters and pre-heaters before
being compressed. The air filters often split into coarse and fine filters trap the dust while the preheaters prevent ice formation which can occur in low temperatures combined with relatively high
humidity at the first blade cascade. The compressed air is then led to the usually circular
combustion chamber, where a small part of the air is mixed with natural gas to form a delated mixture
(for low NOx production) and flows through the burners to the combustion chamber were it is ignited.
The ignited mixture of air/natural gas in the combustion chamber is then mixed with the remaining
compressed air, which enters the combustion chamber through various channels. This produces flue
gases with a temperature of around 1090C for a GE Frame 6, but which can rise to a temperature of
1288C for a modern GE Frame 9FA (applied at EEMS). These high-temperature and high-pressure
gases are led via inlet guide vanes to the rotor blades of the gas turbines expansion turbine, which
then drives the compressor via an axle. The compressor needs power to compress the air from
outside and is connected to the generator by the same axle. Around 2/3 of the power delivered by the
expansion turbine is needed for compression (whereby the temperature of the air rises), the remainder
is for the generator. The gas turbine is essential for the generation of power in a CCGT because,
when unavailable, no power can be produced by the unit.
Heat Recovery Steam Generator (HRSG)
Usually, the exhaust gases from the gas turbine still have a temperature of over 500C. These gases
are lead through a HRSG (see Figure 4.12) and energy is extracted from this gases which are cooled
by transferring the heat, by means of heat exchanges, towards the heating of boiler feed water into
steam. The HRSG is basically a large basin with pipes (parts fitted with fins) in several heat
exchangers banks where boiler feed water flows inside the pipes and the exhaust gases outside the
pipes.

61

Part 2

Figure 4.12: EEMS facility with the exhaust gas boiler on the left (Source: Electrabel)
Condensate water out the condenser, is pumped to the deaerator (a large vessel with boiling water) to
remove the oxygen in the condensate. From there it is called boiler feed water, and this water is
pumped to the feed water pre-heaters and is heated by cooling of the exhaust gases. In the preheater the boiler feed water is heated to a few degrees bellow evaporation temperature and led to the
evaporater in which the boiler feed water is converted into steam. This steam is then overheated in
separate pipe systems (the over-heaters) and led to the steam turbine. To extract the maximum
volume of thermal energy from the exhaust gases different levels of pressure are used to produce
steam see Figure 4.13, and keep the energy losses low. The chosen level of pressure (in the design
phase) depends on the location (whether or not near the gas turbines outlet) in the HRSG and the
temperature of the exhaust gases at that point. If the exhaust gases match the criteria (flow and
temperature) steam can be produced at several levels (usually three levels in larger and modern
facilities, High Pressure (HP), Intermediate-pressure (IP) and Low Pressure (LP)) with or without
reheating the IP steam. At medium-sized facilities production usually takes place at HP and LP. A
system of pumps ensures that the boiler feed water continues to flow through the pipes and that the
energy supply from the exhaust gases to the water remains constant without ebullition. The transfer of
heat in the boiler causes the exhaust gases to cool down to a temperature of around 90C, whereupon
they exit the boiler via a chimney. A HRSG may be fitted with burners for supplementary firing for, say,
extra steam production, but it is called a HRSG because the main source of energy use is the exhaust
gases of the gas turbine.

62

Chapter 4 Energy Supply Chain and Supply Options

Figure 4.13: Typical Q-T diagram of a 3 pressure level HRSG (Source: Kehlhofer et al, 1999)
Steam Turbine
In the larger installations like the EEMS power plant the steam turbine consists of HP, IP and LP
turbines connected by one shaft. The HP steam from the HRSG expands first in the HP turbine to IP
conditions and (in the case of a HRSG with re-heating, like EEMS) is combined with the produced and
not superheated IP steam in the HRSG, whereupon the flows are re-heated together in the HRSG, led
to the steam turbine again and expanding in the IP turbine. The expanded IP steam is then combined
with the LP steam which has been generated in the HRSG and the flows together are expanding in the
LP turbine (see also Figure 4.18) to almost vacuum conditions i.e approximately 0.03 bar and 30C. At
the EEMS power plant the gas turbine and steam turbine have the same shaft and generator.
Compared with a two shaft configuration, with gas turbine and steam turbine both having its own
generator, the one shaft configuration leads to a better overall efficiency. The total energy loss is less
because of the absence of the extra generator and extra generator losses.

63

Part 2

4.5

Combined Heat and Power

The CCGT installations described in the previous section all have condensation steam turbine(s)
without steam extraction. Though expansion is usually maximized in the steam turbine, the steam still
contains a lot of energy (enthalpy) - but this cannot be used because of the almost vacuum conditions
and the low temperature. Hence, the exergy is more or less zero (i.e. no longer capable of generating
power). A 100% condensing operation therefore needs large amounts of cooling water. So, in a large,
modern, CCGT without heat supply and with an electricity output of around 55%, approximately 40%
of the available combustion energy disappears with the cooling water. In other words, low temperature
heat supply can create large fuel savings because far less heat is lost with the cooling water and
alternative sources of heat generation are avoided.
4.5.1

Specific Characteristics of CHP

This study defines combined heat and power (CHP) as follows.


The process whereby usable quantities of heat and electricity are produced simultaneously.
Its great advantage over separate production is that less fuel is needed to generate the same amount
of heat and power, with lower emissions. The general principle of CHP is illustrated in Figure 4.14.
Separate systems

Energy
input

Cogeneration

38
losses

Energy
input

45 pow er

83
Pow er
station
ef f i ci ency
54 %

CHP
ef f i ci ency
8 5%

100

40 heat

44

128 total

Losses

Boiler
ef f i ci ency
90%

15
losses

100 total

4
losses

43 total
Energy saving

15 total
128 - 100
= 22%
128

Figure 4.14: CHP versus separate production and savings compared with reference situation
The most prominent examples of CHP are found in processing industries paper, chemicals, food,
and refining where steam is generated by the process, in glasshouse horticulture for heating
greenhouses, and in the built environment as district heating for heating and hot water.

64

Chapter 4 Energy Supply Chain and Supply Options

With a share of almost 40% of gas-fired CHP generation the Netherlands is one of the countries with
the largest share of CHP. Combined heat and power is therefore vital to the security of supply but
even more important is its potential environmental performance. This section focuses primarily on gasfired CHP and the characteristics of its production process in respect of industrial steam supply and/or
heat distribution (mini and micro CHP are discussed in the Section 6.3.3).
Because this research focuses on electricity generation and because of their limited share in the total,
following forms of CHP are not covered by this research:
1. Gas engines, gas turbines and diesel engines for generating solely mechanical power (for
example driving a gas compressor in compressor stations) from which the exhaust gases are
also used for heat applications.
2. Steam turbines generating solely mechanical power.
4.5.2

Heat Supply Pattern; CHP Back-Up / Peak-Demand Facilities

As a rule, heat supply to an industrial process is a continuous process and to a district heating system
depending on the season. As the main heat source can be unavailable for some reason or another, backup facilities are needed - often complemented by peak-load facilities for heat distribution.
Industrial CHP
Heat supply to an industrial process often takes the form of continuous steam supply throughout the year.
Depending on how the production process is organized there is almost always a back-up facility for the
main heat source. These back-up facilities usually hot-standby steam boilers take over in the case of an
outage of the main heat source. The installed thermal capacity of the boilers is sufficient to cover the
complete steam supply.

100%

35

90%

30

80%

25

70%

20

60%

15

50%

10

40%

30%

20%

-5

10%

-10

0%
0

1000

2000

3000

4000

5000

6000

7000

8000

Heat Load
Amb Temp

Ambient Temperature [C]

Heat Load [%]

Heat Distribution
Extreme sensitivity to temperature and consumption patterns can trigger strong seasonal and daily
fluctuations in the demand for heat (see Figures 4.15 and 4.16).

-15
9000

Time [h]

Figure 4.15: Annual heat supply curve for an average heat distribution project (Source: Own
information)

65

Part 2

100%

90%

80%

70%

Heat Load [%]

60%
Winter

50%

Summer

40%

30%

20%

10%

0%
0

10

12

14

16

18

20

22

24

Hour [h]

Figure 4.16: Load curve for an average heat distribution project over one 24-hour period summer and
winter (Source: Own information)
The best energy results are attained when the entire heat demand is covered by a good source. This is
only possible for projects in which there are multiple heat sources with sufficient capacity. Usually, the
upper part of the demand curve is covered by peak load capacity. This capacity is subject to certain
requirements.
Guaranteed and non-guaranteed heat supply. Guaranteed heat supply needs to be at least as
reliable as the gas grid high-efficiency boiler chain. There are no known in-depth risk analyses
which might account for the well-worn rule of thumb that reserve capacity must be deployed if a
single component ((gas) turbine, boiler) breaks down at peak times ((n-1) criterium). What should
be examined are the consequences of any such breakdown. A gas turbine failure will stop the
CCGT; and a steam turbine failure could be compensated via the steam pressure reducer (HP
steam is reduced to LP conditions suitable for a DH condenser). Back-up facilities are not needed
for the gas, electricity or heat grid because they are already incorporated in the gas grid boiler risk.
However, a trip of one turbine may not lead to the trip of another turbine via electrical connection.
In the case of non-guaranteed supply, the buyer himself is responsible for taking care of the
guarantee. The heat distributor contractually guarantees only a percentage of the supply.
Possible Supply Alternatives
The following alternatives are available:
1. Heat buffers: a distinction needs to be drawn between, on the one hand, physical buffers at
ground level which are intended mainly for optimization, but can also flatten the daily peak and be
deployed in the event of minor disturbances and, on the other hand, large underground buffers
that can absorb seasonal fluctuations and provide full back-up/peak relief. Above-ground buffers
are proven technology, but very little experience has been gained so far in underground buffers.
Buffers that can be realized atmospherically make a considerable difference to the costs
compared with pressurized buffers. The costs of buffers depend also strongly on the situation.
2. Gas boilers, the most commonly used alternative, can be placed near the main heat source or
somewhere in the grid. This is an optimization issue. The main characteristics of gas boilers:
a. The costs must be charged entirely to heat.
b. Deployment is highly limited.

66

Chapter 4 Energy Supply Chain and Supply Options

c.

On the basis of lower heating value the average efficiency from gas boilers in incidental
operation is roughly 85 - 90%.
3. Light Oil or Gas/Light Oil-fired boilers: these are again becoming a realistic option for two reasons:
high capacity and the transport costs for incidental gas consumption. Obviously, this is not a
rational option from an environmental angle.
4. Boilers with bio-fuel: in stead of light oil also bio oil can be applied.
5. Steam from the reducer: this works excellently in installations with a sure supply of steam.
4.5.3

Combined Heat and Power Units for Industrial Heating

With around 123 installations operational (situation 2003), industrial CHP units have the largest share
(approx. 21%) of the installed capacity in the Netherlands. In terms of capacity, these units range from
2 MWe to 475 MWe while using various generation technologies including:
Gas- and steam turbines (CCGT)
Gas turbines (GT) and Heat Recovery Steam Generators (HRSGs)
Boilers and steam turbines (ST)
Gas turbines.
See Figure 4.17 for an overview of the number of units by capacity range and used technology.
130
120
110
100
90
80
Units [-]

TOTAL
CCGT

70

GT + HRSG
GT + FURNACE

60

BOILER + ST
50
40
30
20
10
0
1 - 25

25 - 50

50 - 75

75 - 100

> 100

TOTAL

Installed Power Range [M We]

Figure 4.17: Number of CHP units by capacity range (Source: Cogen, own information)
Of the total 35 CCGT units 9 units have a capacity of more than 100 MWe (see Table 4.9) each and
with 2,057 MWe collectively, they represent around 47% of the industrial CHP CCGT installed power.
Each unit > 100 MWe will be described in more detail in the next section. The industrial CHP share in
installed power of all CCGT amounts to approximately 70% and this indicates that CCGTs are the
most commonly used technology for industrial CHP. The 62 GT+HRSG units are largest in applied
installations. Installed power is for 52 units in the range of 1 - 25 MWe and the remaining between 25 50 MWe. There are 22 Boiler + ST units of which 20 in the range 1-25 MW and the remaining between
25 - 50. The GT+Furnace units complete the list with 4 units.

67

Part 2

CCGT
The above-mentioned 9 larger (> 100 MWe) units supply steam to industry on 8 sites (see Table 4.9).
Electrical efficiency is not mentioned because this is a function of the heat load delivered. The effects
of heat load on electrical, thermal, and total efficiency for two typical CCGT CHP options is given in
Section 8.4.2. The IJmond unit differs from the other units in that it burns the residual gases from
Corus. Below, the characteristics of these larger units and a mid-sized CHP are discussed.

Name or location

Label

CHP Moerdijk
MD-1
Swentibold
SW-1
IJmond
IJM1
Terneuzen Elsta
EL
Delesto 1
DL-1
Delesto 2
DL-2
Salinko Hengelo vof
SA
Yara Sluiskil
HYA
Shell Raffinaderij Pernis PER+ PER

Owner

Commisioning

Essent
1997
Essent
1999
Nuon
1997
Delta / Essent / AES
1998
Essent / Akzo Nobel
1987
Essent / Akzo Nobel
1999
Essent / Akzo Nobel
1994
Yara Sluiskil
1977
Shell Nederland Raffinaderij 1997

Net Power
[MWe]

Pressure
[bar(a)]

Temperature
[C]

339
231
144
475
190
340
104
105
127

100 / IP / LP
HP / IP / LP
65 / 6.5
90
88 / 3.5
104 / 24 / 4
100
82

525 / TIP / TLP


THP / TIP / TLP
405 / 280
510 / 175
537 / 537 / 294
525
525

primary

------ Fuel type -----secondary


tertiary

Gas
Gas
Hgas
Gas
Gas
Gas
Gas
Gas
Gas

Table 4.9: Industrial CCGT units with a capacity of more than 100 MWe (Source: SEP, own
information)
Technology
Applied Cycle
The applied cycles are the Brayton/Joule cycle as topping cycle and the Rankine cycle as bottoming
cycle. The Rankine cycle is implemented as a fired or unfired heat recovery steam generator (HRSG)
behind a Brayton/Joule cycle, the gas turbine. Steam supply for delivery to industrial processes is
extracted from the steam turbine or direct from the HRSG on 1 or more pressure level(s).
Fuel Supply
Natural Gas
The natural gas for the unit is transported under high pressure via the Gasunie Transportservices
(GTS) pipelines to the custody transfer station at the site in which the gas pressure can be adjusted
before the gas is transported to the gas turbine and HRSG burners.
Blast Furnace Gas
Residual gases which are released in the steel production process comprise blast furnace gas, cokes
oven gas and oxy gas. The gas turbine at IJmond 1 is designed for the consumption of mixed blast
furnace gases. As the combustion values for blast furnace gas (4.0 MJ/nm3), oxygas (8.5 MJ/nm3) and
coke oven gas (19.5 MJ/nm3) are lower than for natural gas (31.65 MJ/nm3), large flows of fuel have to
be transported to the gas turbine burners in order to attain the thermal capacity required for
combustion.
Production Process
The CCGT installations consist of gas turbine(s), heat recovery steam generator(s) unfired or
supplementary fired, and back-pressure or condensing steam turbine(s) with steam extraction. Without
the steam extraction, the production process of the gas turbine(s) and the heat recovery steam
generator(s) is identical to the production process of the Combined Cycle Gas Turbine (CCGT).
Steam Turbine with Steam Extraction
In the larger installations the steam turbine consists of high-pressure, intermediate-pressure and lowpressure turbines connected by a shaft. First, the high-pressure steam from the HRSG expands in the
high-pressure turbine to intermediate-pressure status (i.e. in the case of a HRSG with re-heating)
whereupon it is merged with the intermediate-pressure steam. The flows are then re-heated together
in the HRSG before expanding in the intermediate-pressure turbine. The expanded intermediatepressure steam is then merged with the low-pressure steam which has been generated in the HRSG
before expanding in the low-pressure turbine (see Figure 4.18).

68

Chapter 4 Energy Supply Chain and Supply Options

Figure 4.18: Process scheme of Delesto 2, triple-pressure HRSG with reheat and steam turbine with
steam extraction (Source: Akzo Nobel et al., 1996)
Steam can be extracted from the turbine at one or more levels of pressure depending on the needs of
the process. In most units steam extraction takes place at only one pressure (e.g. low pressure in the
case of Delesto 2). At most of the units steam is extracted by means of a control valve mechanism to
ensure a constant level of steam pressure to the process. The remaining expanded steam leaves the
turbine under very low pressure (almost vacuum conditions of approximately 0.03 bar) and at a low
temperature (approximately 30C) in order to bring about a maximum drop in enthalpy and hence
maximum output of the steam turbine. The expanded steam is condensed to water in the cold
condenser and returned to the heat recovery steam generator by condensate pumps.
The process in mid-sized and smaller CCGTs with heat supply, though very similar to the process in
larger units, is far less complicated because, instead of three levels of pressure in the HRSG, only two
are used or even one and also reheat in the HRSG is commonly not applied. Therefore HRSG and
steam turbines are less complex (see Figure 4.19).

Figure 4.19: Process scheme of CHP Schoonmansmolen, a double-pressure HRSG and steam
turbine with steam extraction (Source: Stork Ketels, 1995)

69

Part 2

CHP Moerdijk
The CHP Moerdijk is located on the industry terrain Moerdijk adjacent to a large municipal waste
incineration plant, AVI Moerdijk (3), see Figure 4.20. The CHP plant is an example of process
optimizations since the CHP generates not only electricity by use of natural gas but also by the use of
280 t/h high pressure (100 bar) steam imported form the waste incinerator (1). At the same time it
provides Shell Nederland Chemie with 145 t/h of intermediate pressure steam (6).

Figure 4.20: Overview of the fully integrated AVI and CHP plant at Moerdijk (Source: Essent, 2001)
The CHP plant at Moerdijk has three gas turbines, each with its own triple-pressure HRSG with
intermediate-pressure re-heating, and has one communal condensing steam turbine with steam
extraction. Each turbine has its own generator. Shell gets its steam from the intermediate-pressure
turbine. Steam from the waste incineration plant enters at a pressure of 100 bar and a temperature of
around 400C and is heated in the HRSG to 525C overheated steam. Water from a natural draft
cooling tower with a height of 103 meters is used as cooling water to cool the condenser with a view to
the remaining expanded steam. Natural draft cooling towers exploit the thermo-dynamic behavior of
fine-misted water droplets, which need thermal energy in order to evaporate. This process cools the
cooling water. The shape of the cooling tower (wide at the bottom (78 meters) and narrow (48 meters)
at the top) creates the natural draft that is needed to remove the water vapor with displaced air. As a
consequence of coupling the CHP plant with the waste incinerator, municipal waste is converted into
power at a efficiency of 28% (Fris et al, 2003).
Swentibold
The CHP plant Swentibold has one gas turbine, two triple-pressure HRSGs, and one steam turbine
with steam extraction for supplying process steam, see Figure 4.21. It forms the new technological
core of the utility services for the North and South sites of DSM in Geleen (the Netherlands) and
supplies the DSM factories with steam at various levels of pressure. To optimize security of supply the
entire system is designed in such a way that steam can still be delivered if one component is nonoperational. It is with this in mind that the plant has been equipped with two HRSGs - each with
supplementary firing - suitable for natural gas and/or residual gas from DSM installations - and
combustion air ventilators in case the gas turbine breaks down. Swentibold also has a natural draft
cooling tower.

70

Chapter 4 Energy Supply Chain and Supply Options

Figure 4.21: Process scheme of Swentibold, triple-pressure HRSG and steam turbine with steam
extraction (Source: Essent, 2005)
IJmond
IJmond is a CHP plant with one gas turbine, one double-pressure HRSG, and one condensing steam
turbine with steam extraction. It supplies Corus with process steam. Like the Velsen units, IJmond is
situated on the Corus terrain. What distinguishes the plant is the fact that it is fired by three types of
low-calorific blast furnace gas which is enriched with natural gas in a gas turbine specifically designed
for this purpose. The turbine, generator, steam turbine and gas compressor (for compressing the blast
furnace gas to the right pressure for combustion in the combustion chamber) are connected by one
shaft. The gas compressor is connected to the shaft by a gearbox. Using steam from the Corus grid
the steam turbine serves as a start turbine during the start-up of the installation.
Elsta
Elsta is a CHP plant which is powered by British gas and Dow sitegas. The British gas has been
entering the Netherlands via the interconnector since October 1998. The pipeline branches off at
Zelzate (Belgium) and the gas is then transported via the ZEBRA pipeline (an Essent/Delta joint
venture) to the Elsta CHP plant in Hoek. Sitegas which Elsta uses in addition to British gas is
released when LPG is cracked. It comes from Dow and is not available all the time. The quantities are
also subject to variation. The gas is transported via a pipeline from Dow to Elsta.
The installation (see Figure 4.22) consists of three GE-type 9E gas-fired turbines, each with an
approximate capacity of 123 MWe, and a GE steam turbine with an electrical capacity of
approximately 92 MWe at full load. Each of these turbines drives a generator which produces
electricity. The maximum electrical capacity that can be generated is approximately 452 MW. A HRSG
is connected to each gas turbine. Together they generate high-pressure steam (850 t/h of 90 bar) and
low-pressure steam (7 bar). The 90 bars of steam are partially reduced to 35 bar via the steam turbine
and/or pressure reduction valves and delivered to the nearby chemical plant of Dow Benelux. The
HRSGs are equipped with supplementary firing to further increase the steam output, if necessary.

71

Part 2

On average, 500 t/h steam is delivered to Dow. The remainder expands in the steam turbine until it
reaches almost vacuum conditions and is condensed in cooling towers fitted with ventilators. It is then
re-used as feedwater for the HRSGs.

Figure 4.22: Overview of the Elsta CHP plant (Source: Elstacogen)


Delesto 1
At present there are two Akzo Nobel divisions in Delfzijl (Akzo Nobel Chemicals bv and Akzo Nobel
Salt) and three Joint Ventures (Delamine bv, Methanor vof and Delesto bv). Delesto is the energy
center, generating both heat and power for the Delfzijl site. Delesto currently consists of Delesto 1 and
Delesto 2 and has had CHP generation units since as far back as 1957.
The first large CHP plant of Akzo Nobel in Delfzijl (Delesto 1, see Figure 4.23) became operational in
1987. It consists of three GE-type Frame 6E gas turbines, each with the capacity to generate
approximately 37 MWe of electricity. Connected to each gas turbine is a double-pressure HRSG
where steam is generated of 88 bar and 3.5 bar. The HRSGs are equipped with supplementary firing
meaning the steam production in each case can range from 80 ton/h (62 t/h at 88 bar and 18 t/h at 3.5
bar) without supplementary firing to 160 t/h (all at 88 bar) with maximum supplementary firing. The
supplementary firing facility is suitable for both natural gas and hydrogen gas. The hydrogen gas is a
by-product of the electrolysis process in which electricity is used to separate salt (sodium chloride) into
sodium (dissolved in mercury to sodium amalgam) and chlorine gas. When the sodium amalgam is
brought into contact with water it results in sodium hydroxide and hydrogen gas. The mercury is
returned to the electrolysis cell and the hydrogen gas is dried and compressed to a pressure of 8 bar
before being delivered to the exhaust gas boilers. The steam generated in these HRSGs goes to two
back-pressure steam turbines with steam extraction at 30 bar for the delivery of process steam and
3.5 bar for the delivery of process steam and heating.

72

Chapter 4 Energy Supply Chain and Supply Options

Figure 4.23: Process scheme of the CHP system at Delesto 1 (Source: SBB, 1990)
Delesto 1 produces 166 MWe at a maximum steam output of 480 t/h with a total efficiency of 89%.
However, it is usually deployed for the delivery of 360 t/h in steam whereby around 152 MWe is
delivered. A GE-type Frame 5 unit with a HRSG can be used as a back-up unit when, for example,
one of the gas turbines is being serviced.
Delesto 2
Delesto 2 is the second CHP plant of Akzo Nobel in Delfzijl. Together with Delesto 1, it is a highly
flexible operation. Delesto 2 delivers 340 MWe and has a steam output varying from 0 to a maximum
of 320 t/h. With no steam production, Delesto 2 is a conventional CCGT with a net electrical efficiency
of approximately 55% at full load; but with maximum steam production, total efficiency can rise to
around 81% (see Figure 8.12 of Chapter 8).
The CCGT is a single-shaft unit in which all the components, from the gas turbine to the generator, are
connected in succession (see Figure 4.18).
The gas turbine is a GE-type Frame 9FA, similar to the gas turbines at the EEMS plant (see Section
4.4.7). It is linked to a triple-pressure HRSG with intermediate-pressure re-heating, feeding a
condensing steam turbine with controlled steam extraction in the low-pressure section. The exhausted
expanded steam is condensed in a seawater-cooled condenser and is re-used as feedwater for the
HRSG.
Salinco
In Hengelo (Dutch province of Overijssel) also two Akzo Nobel divisions (Akzo Nobel Chemicals bv
and Akzo Nobel Salt) are located and one joint venture (Salinco). Salinco is the energy center for the
Hengelo site and produces heat and electricity. The original CHP installation consists of two highpressure steam boilers and two back-pressure steam turbines with steam extraction at two pressure
levels. In 1994 a new installation became operational, consisting of a GE-type Frame 6B gas turbine
and a HRSG with supplementary firing facilities which can deliver up to 175 t/h of steam when
operating at maximum capacity. As in Delfzijl, the supplementary firing is suitable for hydrogen gas.

73

Part 2

The expansion of the generating facilities has led to one high-pressure steam boiler being used as a
standby. Salinco delivers 76 MWe for a heat requirement of 137 MWth.
Yara (formerly Hydro Agri)
The Yara CCGT installation consists of a gas turbine, HRSG with supplementary firing, cold standby
boiler and 2 steam turbines. Together the installed power is approximately 105 MW and dates from
1977. Rebuild of the plant is taking place or will take place in the near future in which a new gas
turbine is installed and after which the second steam turbine will become a cold standby facility.
PER+
The PER+ CHP plant is exceptional for a number of reasons, the main being deep integration of the
CHP process with the refinery at Pernis (see Figure 4.24). This integration involved more than the
delivery of steam to the refinery grid; it also included the use of residual gases from other processes in
the plant, such as syngas, which is produced in the gasification installation. The integration has been
designed in such a way that the gasification installation cannot operate stand-alone, the plant however
can. The major issue at the plant is therefore security of supply, followed by emission reduction and
optimal output. The plant consists of a number of installations constructed around a core comprising
two General Electric Frame 6 (2 x 43 MWe with syngas as fuel) gas turbine generators in combination
with supplementary fired natural-circulation single-pressure HRSGs with a maximum steam capacity
of 238 t/h high-pressure (82 - 88 bar(g), 525C) steam. 124 t/h of this is saturated steam which is
imported from the gasification unit and fed direct into the steam drums of both HRSGs. Through a
combination of reducers, a back-pressure steam turbine and a condensing steam turbine the highpressure steam is reduced and three steam grids (high-pressure, intermediate-pressure, lowpressure) of Shell Pernis are fed while electricity is generated at the same time.

Figure 4.24: Overview of the PER+ refinery at Pernis, from the left to the right : Hydrocracking Unit, Oil
gasification installation (residual gasifier) and the CHP Unit (Source: Energietechniek, 1997)
CHP < 100 MWe
The capacities of the other 26 CCGTs varies from 6 MWe (VUmc) to approximately 90 MWe
(Eurogen), most being in the range of 25 - 50 MWe. The difference between these units and the larger
CCGTs is attributable primarily to the lower heat demand at the location where the unit is situated.
This, in combination with the choice for an optimal heat-power ratio, has prompted decisions for
smaller gas turbine(s) and a simpler steam cycle with, for example, a double-pressure or singlepressure HRSG with supplementary firing facilities instead of a triple-pressure HRSG with re-heat and

74

Chapter 4 Energy Supply Chain and Supply Options

required steam turbine with triple pressure admission and steam extraction. The electricity production
is largely determined by the gas turbine, which can be deployed in partial load. Usually the
supplementary firing can also be used to produce extra steam which can be converted into extra
electricity in a condensing steam turbine. The deployment options for a larger CCGT with one Frame
9FA and a mid-sized CCGT with two GT10s are shown in Figures 8.12 and 8.14 of Chapter 8.
The deployment options are particularly dependent on the heat demand. If this is high, the options are
limited; there is more scope for deployment when the heat demand is low.
GT + HRSG
The GT + HRSG installations consist of a gas turbine and usually a supplementary fired singlepressure HRSG (see Figure 4.25).

Figure 4.25: Process scheme Hunzestroom GT + HRSG plant (Source: Energietechniek, 1998)
Normally, the gas turbines have a capacity of up to 40 - 50 MWe. The HRSG are fairly basic in design
and consist in effect of only an economizer section, a vaporizer section and, if necessary, an
overheater section plus supplementary firing burners to absorb fluctuations in the demand for steam.
These fluctuations can also be leveled out with a conventional steam boiler, provided one is available
and the economics allow it. The electricity production is determined entirely by the gas turbine, which
can be deployed in partial load. Needless to say, this lowers the steam output, which must then be
raised again by extra supplementary firing in the HRSG or conventional boiler (see Figure 8.16 and
8.18 of Chapter 8).
Boiler + ST
Most of the boiler + ST installations are older installations from the 1960s and 1970s, in which a
conventional steam boiler for producing high-pressure steam is connected with a (back-pressure)
steam turbine where the high-pressure steam from the boiler expands to the required end-pressure.
The heat-power ratio tips in favor of heat (see Figure 8.20 of Chapter 8) as the share of electricity is
very limited. The electricity production is determined entirely by the demand for heat. There are no
other deployment options (apart from orthodox steps such as releasing steam or bypassing the steam
turbine).
GT + Furnace
GT + Furnace are typical applications of gas turbines as hot air ventilators for, e.g. the ovens of crude
distillers at refineries. The number is limited and the gas turbines are used, if available, as this means
that primary oven fuel can be avoided and additionally electricity can be generated.

75

Part 2

4.5.4

Combined Cycle Gas Turbine District Heating (CCGT DH and GT+HRSG DH)

Present-day, 12 larger steam and gas turbine (CCGT) units and 2 smaller gas turbine + HRSG units
(ROC1 and ROC2) distributed over nine sites supply heat to a heat distribution grid in the
Netherlands (see Table 4.10). The heat distribution projects in question are: Almere, Leiden, Den
Haag, Rotterdam, Utrecht, Amsterdam, and Purmerend.

Name or location

CHP Almere
CHP Almere
Leiden
Den Haag
R'dam Galileistr
RoCa
RoCa
RoCa
Lage Weide
Merwedekanaal
Merwedekanaal
Merwedekanaal
Diemen
Purmerend

Label

AL-1
AL-2
LD12
GV15
FG-1
ROC1
ROC2
ROC3
LWE6
MK10
MK11
MK12
DM33
PU-1

Owner

Electrabel
Electrabel
E.ON
E.ON
E.ON
E.ON
E.ON
E.ON
Nuon
Nuon
Nuon
Nuon
Nuon
Nuon

Commisioning

1987
1993
1986
1982
1988
1982
1982
1996
1995
1978
1984
1989
1995
1989

Net Power
[MWe]

Max Th Power
[MWth]

Pressure
[levels]

Temperature
[levels]

64
54
81
78
209
24
24
221
247
96
103
224
249
69

64
54
83
76
256
40
40
200
180
100
110
180
180
65

HP / LP
HP / LP
HP
HP
HP
Hot Water
Hot Water
HP / IP / LP
HP / IP / LP
HP
HP
HP / IP / LP
HP / IP / LP
HP / LP

THP / TLP
THP / TLP
THP
THP
THP

THP
THP
THP
THP
THP
THP
THP

/ TIP / TLP
/ TIP / TLP

/ TIP / TLP
/ TIP / TLP
/ TLP

primary

------ Fuel type -----secondary


tertiary

Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas
Gas

Table 4.10: Heat distribution units2 (Source: SEP, own information)


Small-scale heat distribution projects are located in Enschede, Erica/Klazinaveen and Helmond, of
which the Erica/Klazinaveen project supplies heat to greenhouses. The technology is similar to a
larger CCGT DH but as capacity is geared to the basic-load heat demand in those areas these units
are smaller in terms of electrical capacity.
Technology
Applied Cycle
The applied cycles are the Brayton/Joule cycle as topping cycle and the Rankine cycle as bottoming
cycle. The Rankine cycle is implemented as a fired or unfired heat recovery steam generator (HRSG)
behind a Brayton/Joule cycle, the gas turbine. Steam supply for district heating is extracted from the
steam turbine. In some units DH Coils bundles at the end of the HRSG provide the waste heat part of
the heat delivery.
Fuel Supply
Natural Gas
The natural gas for the unit is transported under high pressure via the Gasunie Transportservices
(GTS) pipelines to the custody transfer station at the site in which the gas pressure can be adjusted
before the gas is transported to the gas turbine and HRSG burners.
Production Process
The CCGT DHs process is well comparable with a CCGT, the few differences being caused mainly by
the heat supply and the units age.
Heat Recovery Steam Generator
Whereas all conventional CCGTs (see Section 4.4.7), except for the Donge power station, have a
HRSGs with three pressure levels and IP reheating facilities, only the Roca (ROC3), Diemen (DM33)
and Lage Weide (LWE6) CCGT DH units have the same layout. The other units have either a HRSG
with two pressure levels (AL1, AL2, MK12, PU1), with MK12 having reheating facilities for the IP
steam from the steam turbine, one pressure level (LD12, GV15, FG-1, MK10, MK11), with FG-1
2

The mentioned maximum thermal power in Table 4.10 is only the CCGT thermal power. This thermal power
normally covers 1/3 of the maximum heat load (MWth) of a heat distribution project. The remaining 2/3 is covered
by gas/oil boilers. The 1/3 part of the CCGT however is enough to cover 80 - 90% of the total required heat (GJs)
of the heat distribution project.

76

Chapter 4 Energy Supply Chain and Supply Options

having supplementary firing facilities for additional steam generation, or just a hot water component
(ROC1, ROC2). Except for LWE6, DM33 and MK12 all units have a separate district heating bundle in
the upper part of the HRSG. One distinctive feature of the ROC3 unit is that some of the cooled
exhaust gases are led to a CO2 boiler, where sophisticated combustion technology is applied to
supplementary fire extra natural gas in order to maximize the CO2 concentration in the flue gases.
These enriched flue gases are then compressed and transported to the horticulturists via a special
network. The extra steam produced is led to the steam turbine.
Steam Turbine
A typical steam turbine for CCGT DH purpose works in more or less the same way as in a combined
cycle gas turbine (CCGT), whereby the number of feed-in points varies according to the pressure
levels in the HRSG. The difference is that, in CCGTs which are used for District Heating, steam is
extracted for heat production in DH condensers this is in contrast with standard CCGTs where all
the steam expands until it reaches almost vacuum conditions. The steam extraction takes place at
pressures of between 0.8 and 4 bar and depends on the number of extraction points in the steam
turbine. Thus, if there are two extraction points, one extraction will be at a relatively low pressure of,
say, 0.8 - 1.5 bar and the other at a slightly higher pressure of 3 - 4 bar. The latter extraction takes
place, amongst others, when the demand for heat exceeds the capacity of the first extraction and DH
condenser. It is also used to get the required supply temperature if it is higher than 100C. At AL1,
AL2 and PU, the expanded steam, instead of only being extracted for DH condensers (see steam
turbine part in Figure 4.26), is cooled in a condenser which uses the DH return as a cooling medium.
As for supply temperature reasons also an extraction point is available.

Figure 4.26: CHP plant and district heating system in Almere (Source: EPON, 1980s)
This leads to the best overall efficiency; however, it also means that if the demand for heat is low and
hence the cooling capacity of the DH return is also low (e.g. summer) then deployment of the unit is
limited. This problem is largely solved by heat buffers which can store heat during the day and supply
it at night. That way, the unit can be taken out of operation during the night. Heat buffers are used in
Almere, Purmerend, The Hague and Rotterdam amongst others.

77

Part 2

Buffers
But this is not the only purpose that heat buffers are used for (see Figures 4.27). For example:
Top-ups to the electricity supply during workdays (peak) come from heat in buffers which are
filled during the night.
If enough heat can be stored in the buffer, for example, to get through the weekend, there is
no need for a must run at certain times of the year (e.g. summer).
Buffer deployment lowers investment in peak boiler capacity, as a buffer can supply a constant
flow of heat to a DH grid for several hours.

Figure 4.27: Heat buffers, each 4000 m3 at the CHP plant in Purmerend (Source: UNA, 1996)

78

Chapter 4 Energy Supply Chain and Supply Options

4.5.5

Small-Scale Combined Heat and Power (Gas Engines)

At the end of 2003 the cluster of small-scale CHP (gas engine) capacity accounted for approximately
1513 MWe of installed capacity distributed across various sectors (see Figure 4.28).
1,800
1,600

Installed Capacity [MWe]

1,400
1,200

1998
1999

1,000

2000
2001

800

2002
2003

600

2004
2005

400
200

Other Producers

Waste Incineration

Healthcare

Distributors

Other Industry

Chemical Industry

Paper Industry

Food & Luxury Goods

Refineries & Mining

Agriculture &
Horticulture

Total Small-Scale
CHP

Figure 4.28: Installed small-scale CHP (gas engines) capacity distributed across various sectors
(Source: CBS statline)
Agriculture & Horticulture has by far the highest share in the total small scale combined cycles. In
2003 it amounts to 960 MWe on a total of 1,513 MWe, being a share of approximately 63% in 2003
and this has grown to a share of 67% (1,120 MWe on a total of 1,675 MWe) in 2005. After an initial
rise between 1998 and 2000, the total installed capacity declined in 2001, 2002, and 2003 and then
took an upward turn (see Figure 4.28).
The capacity in the horticultural sector follows a particularly interesting pattern of development. After
an initial decline in the overall installed capacity, the share of the distribution companies is declining
still further while the share of the horticulturists is rising (see Figure 4.29). This is because of the
transfer of installations from distribution companies to horticulturists, and the steady increase in
lighting (1,910 hectares were lit in the horticultural sector in 2003; this figure is expected to rise to
3,900 - 5,000 hectares by 2010 (Cogen Projects, 2004)). It is present-day more economical per kWh
for a horticulturist to have a fine-tuned CHP installation with minimum or no heat surplus than to rely
entirely on power from the public grid. An optimal balance still, however, needs to be struck between
the generation of electricity, the production of heat, the purchase of electricity and whether or not
the deployment of flue gas cleaning and heat buffers.
Though smaller gas engines of 300 - 400 kWe are still by far the most popular (CBS / Cogen Projects,
2004), many new installations only use gas engines with a capacity of more than 1 MWe. Clustered
projects or horticultural areas create bigger markets for the energy products and make larger
installations feasible.

79

Part 2

800
Distributors

Horticulture

1-1-2002

1-1-2003

700

Installed Capacity [MWe]

600

500

400

300

200

100

0
1-1-1999

1-1-2000

1-1-2001

1-1-2004

1-1-2005

Figure 4.29: Trend in installed small-scale CHP capacity in the horticultural sector (Source: Cogen
Projects, 2005)
Technology
Applied Cycle
The applied cycle for the small-scale CHP (gas engine) is a Otto/Rankine cycle. The bottoming cycle
is in all units implemented as a waste heat recovery generator producing mostly hot water.
Fuel
Natural Gas
The natural gas is transported under high pressure via the Gasunie Transportservices (GTS) pipelines
to the custody transfer stations in which the gas pressure can be adjusted before the gas is
transported through the pipelines of the regional grid operator to horticultural businesses, healthcare
organizations, and other customers. The grid pressure of this gas is usually between 1 and 8 bar.
Bio Gas
See Section 4.6.4 on biomass fermentation.
Production Process
A gas engine is derived from a conventional cylinder-based diesel engine with entrance and exit
valves for letting in fuel and air and discharging flue gases. The fuel is not, however, diesel but natural
gas or even biogas, propane gas, or other alternatives.
The capacity for, say, Jenbacher gas engines running on natural gas with NOx < 250 mg/Nm3 varies
from 312 kWe (Type 208) to 3,047 kWe (Type 620). The capacity of the gas engine is determined by
the number of cylinders, cylinder content, the fuel, the rpm, and the NOx emissions. Table 4.11 shows
a manufacturers overview, including the electrical and thermal efficiencies.

80

Chapter 4 Energy Supply Chain and Supply Options

Table 4.11: Overview of GE Jenbacher Type 4 gas engines (Source: GE Jenbacher, 2005)
A small gas-engine-based CHP installation has heat exchangers for the recovery of heat as well as
the bare engine. Heat is recovered from:
the flue gases
the intercooler
the lubrication system.
The temperature of the heat recovered from the flue gases is relatively high (e.g. 140C) while the
temperature of the heat recovered from the intercooler and the lubrication system is lower. Total
efficiencies of over 90% can be realized if this heat is used productively.
Flue Gas Cleaning
When natural gas is combusted in gas engines, emissions such as NOx, CH4, and C2H4 (ethane,
ethylene) are released. These emissions are so highly concentrated that the flue gases cannot be
used directly for CO2 dosing (to promote plant growth) in the horticultural sector. Flue gas cleaning
systems can be applied to reduce the concentrations, thereby enabling the exhaust gases from a gas
engine to be used for CO2 dosing. These systems are based on a selective catalytic reduction (SCR)
catalyser with an urea injection in combination with an oxidation catalyser. To avoid crop damage they
are subject to extremely stringent criteria. This explains why the range is still limited. These systems
avoid CO2 fertilization from gas boilers, which is widely applied because the flue gases can be used
immediately hence without cleaning on greenhouse plants.
Flue gas cleaning has the following advantages:
the operational time is higher, making the CHP installation more profitable;
higher energy savings in the greenhouses because boiler for CO2 production can be avoided.

81

Part 2

4.6

Sustainable Options

Renewable energy is a fairly broad term. Besides sustainable generated electricity and imported
green electricity, it covers solar/thermal energy, heat pumps, heat/cold storage, and industrial and
domestic wood-burning stoves expressed in saved primary energy. This research addresses only the
national production of sustainable electricity.
In 2003 national production of electricity from sustainable sources accounted for 3.3% of the electricity
consumption in the Netherlands, see Table 9.9 of Chapter 9. Sustainable electricity in the Netherlands
is generated from:
Wind energy
Hydro-energy
Photo-voltaic solar energy
Bio-energy
o
Biomass incineration
Centralized co firing
Local co firing
o
Biomass fermentation
Landfill gas
Waste water treatment
Others.
o
Waste incineration
4.6.1

Wind Energy

Wind energy is developing fast in the Netherlands, see Figure 4.30. Whereas in 1998 1193 wind
turbine installations were placed, in 2005 this increased to 1709 installations (CBS, 2007).

1500

1224
1250
1073

906

Installed Capacity [MWe]

1000

670

750

485
410

500

447

363

250

0
1998

1999

2000

2001

2002

2003

2004

Figure 4.30 : Development of windcapacity in the Netherlands (Source: CBS Statline)

82

2005

Chapter 4 Energy Supply Chain and Supply Options

Locations of onshore wind parks are usually locations in wind rich areas like close to the North Sea,
Wadden Sea or IJssel meer, see Figure 4.31 in order to catch as many wind as possible which leads
to a higher capacity factors compared with land inward locations. The typical capacity factor in 2003,
2004 and 2005 was approximately 20% (CBS, 2007) which is equal to 1,750 full load hours.

Figure 4.31 : Locations of windparks in the Nethgerlands (Source: Wind Service Holland)

4.6.2

Hydro Power

Hydro-power installation in the Netherlands can be seen as Small Hydro in international context
(ECN, 2003). The installations described in this section are non-pumped-storage hydro plants. There
are at present three large hydro-power installations at distribution companies with a capacity of
between 10 and 15 MWe, see Table 4.12.

83

Part 2

Name or location

Label

Hagestein
Maurik
Linne
Alphen/Lith
Vecht/De Haandrik
Roermond
Particulieren

Commisioning

Net Power
[MWe]

1958
1988
1989
1990
1988
2000
-

1.8
10
11.5
14
0.1
0.2
0.1

Table 4.12: Hydro-power installations (Source: ECN, 2003, own information)


Including the small installations the total capacity of hydro-power in the Netherlands is approximately
37 MWe (CBS, 2007). To generate hydro-power, turbines are used to convert the very weak fall in
Dutch rivers into electricity. The average capacity factor for 2003, 2004, and 2005 was approximately
26% (CBS, 2007) which is equal to 2,275 full load hours.
.
4.6.3

Photo-Voltaic Solar Energy

There are three systems of photo-voltaic solar energy:


autonomous systems
centralized grid-linked systems
local grid-linked systems.
Autonomous systems are not connected to the grid and are applied in areas where there are no
connections to the electricity grid, such as recreational sites for summer houses, caravans, yachts etc,
and in various professional fields such as shipping (solar-powered buoys and beacons) and
agriculture/horticulture (drinking troughs on grazing land).
Most centralized grid-linked systems consist of PV panels mounted on masts which are erected on the
users premises, or covering large surfaces (Floriade).
Local grid-linked systems are applied to buildings (e.g. housing), amongst others. In 1998 there is a
total of approximately 6 MWe(p) in installed photo-voltaic systems and this increased to 46 MWe(p) in
2003 and 51 MWe(p) in 2005. By far the largest share, almost 39 MWe in 2003 and 43 MWe in 2005,
is in the form of local grid-linked systems.
The number of full load hours represents the number of notional hours that photo-voltaic solar energy
systems deliver peak load, resulting in the annual production. The typical full load hours in 2003, 2004
and 2005 were approximately 660 h representing a capacity factor of approximately 8% (CBS, 2007).
4.6.4

Bio-Energy

Biomass Incineration

Central co-firing
This concerns the co-firing of biomass in power plants; for example, wood in coal-fired
power plants (Gelderland 13, Borssele 12, Amer 9). It is achieved by mixing the wood
with the main fuel or by wood gasification (Amer 9). The Claus and Harculo 60 power
plants can co-fire bio-oil.

Local
This concerns local generation of electricity in smaller incineration plants where biomass
is the only fuel. The biomass (wood prunings, waste wood) is converted into thermal
energy by incineration on a grate. The released heat then goes to a steam boiler where it

84

Chapter 4 Energy Supply Chain and Supply Options

is converted into steam which drives a turbine to produce the main product, electricity.
There are at present four biomass plants in the Netherlands with a combined installed
capacity of around 27 MWe. The principal unit is the biomass plant in Cuijk with a
capacity of approximately 25 MWe.

Biomass Fermentation

Landfill gas
Landfill gas is biogas with high levels of methane which is formed by the biological
decomposition of dumped (usually covered) waste. If available in sufficient quantities,
landfill gas is converted into electricity and possibly into heat by combustion in a gas
engine, and can be burnt off if it is less usable because of the decrease in the methane
content. Because no waste dump takes place anymore the amount of this type of biogas
will decline in the future.

Waste water treatment


In waste water processes biogas is extracted via anaerobic treatment. This biogas is
used to generate electricity and heat, or is used directly as gas.

Others
This covers mainly biogas that is extracted and used in the food and the paper industries.
The biogas is extracted through anaerobic waste water treatment and is used to generate
electricity and/or process heat. Biogas is also extracted from organic compost and
manure.

Biomass Others
These are the not by incineration and fermentation covered projects like projects which
use bio oil in for bio oil use modified diesel engines or the use of bio oil in boilers from
which the steam is used to drive a steam turbine.

Waste Incineration

The Netherlands produces 58 million tons of waste every year, 79% of which (46 million tons) are recycled.
Most of this consists of glass, wood, paper, metal, minor chemical waste, and textiles. The other 12 million
tons (21% of the total) are either converted into compost, incinerated or dumped. At present some 5 million
tons are incinerated, 6.5 million tons are dumped, and the rest becomes compost. Out of the 6.5 million
tons of dumped waste, 1.4 million are combustible. These should be incinerated but there are not enough
waste incineration plants (HVC, 2006).
The main purpose of waste incineration plants is to process waste. So, if technically possible, waste
incineration plants almost always run at full capacity. Provided there is good legislation and regulations, the
continuity of waste incinerators is guaranteed. After all, chances are that garbage will always be around. As
waste incineration plants are subject to stringent emission regulations, it is unlikely that they will be
marginalized by other installations. Neither will they switch to clean biomass on a large scale. The electricity
yield from waste incineration installations is still low at an average of only 25%. This is mainly caused by
the high own electricity consumption for flue gas cleaning purposes and the applied steam pressure and
temperatures which are significant lower ( 40 bar, 400C) compared with conventional coal and gas fired
units. However, this means that there is a large heat potential, and that higher energy efficiency can be
achieved with heat utilization than with electricity production alone. Many waste incinerators are able to
utilize heat without lowering the electricity production; the heat from grate cooling could, for example, be
utilized.
In the Netherlands there are currently 13 incineration installations which are collectively able to
incinerate over 5.1 million tons of waste a year (see Table 4.13) and which have an installed electrical
capacity of approximately 414 MWe (CBS, 2006).

85

Part 2

Name or location

Province

Town

Commisioning

Net Power
[MWe]

Essent Milieu GAVI


Drenthe
Wijster
54
1996
Twence
Overijssel
Hengelo
30
1997
ARN
Gelderland
Weurt
28
1987
AVR Avira
Gelderland
Duiven
31.4
1975
Afval Energie Bedrijf
Noord-Holland
Amsterdam
80
1993
Huisvuilcentrale N-H
Noord-Holland
Alkmaar
51.5
1996
AVR Rijnmond
Zuid-Holland
Rozenburg
108
1973
AVR Chemie DTO
Zuid-Holland
Rozenburg
AVR Rotterdam
Zuid-Holland
Rotterdam
25.5
1963
Gevudo
Noord-Brabant
Dordrecht
9.9
1973
ZAVIN
Noord-Brabant
Dordrecht
1991
AZN
Noord-Brabant
Moerdijk
1997
SITA ReEnergy
Noord-Brabant
Roosendaal
1976
Table 4.13 : Waste Incineration installations (Source: Vereniging Afvalbedrijven, 2004; CBS, 2006;
own information)

Most of the waste is domestic, but some industrial waste is also processed. The primary task of a
waste incineration plant is to process (partially recycle) and incinerate domestic and industrial waste in
a responsible manner. The waste is transferred with a grab to a bunker and led to the furnace via
filters. The waste is then incinerated on a sliding grate at a temperature of between 800 and 1000C.
The heat that is released by the incineration process is reclaimed in a boiler where water is converted
into steam. This steam is then used to drive a turbine which generates electricity as the main product
and heat as a by-product. The flue gases are cleaned in a three-step procedure: removing fly ash, wet
washing to remove acidifying components, and finally a DeNOx installation (see Figure 4.32).

Figure 4.32: Schematic overview of a waste incineration plant (Source: HVC)


Waste incineration plants are good sources for process heat and heat distribution projects, provided
sufficient concentrated heat demand is available. Almost all waste incineration plants in the
Netherlands are situated in densely populated and/or industrial areas.
Seven waste incineration plants deliver heat: SITA ReEnergy, ARN Nijmegen, AZN Moerdijk, AVR
Rozenburg, AVR Avira Duiven, HVC Alkmaar, and AEB Amsterdam. SITA ReEnergy for drying sludge
and heating greenhouses, ARN in the form of low-grade heat for the Rivierenland sewage treatment
plant, AVR Rozenburg for the production of distilled water and the Kerr-McGee production process
(Energietechniek, 2004), AZN Moerdijk in the form of intermediate pressure steam for the nearby CHP
plant at Moerdijk (see Section 4.5.3), and AVR Avira, HVC and AEB for city heat distribution. Around
50% of the (domestic) waste that is used to fire Dutch waste incineration plants is biogene, so almost
50% of the generated energy can be classified as renewable (Vereniging Afvalbedrijven, 2004).

86

Chapter 4 Energy Supply Chain and Supply Options

4.7

Technical Conditions for Unit Deployment

Heat Consumption Curves and technology development over the years


Figure 4.33 displays the specific heat consumption curves (i.e inverse electrical efficiency: 3600/e) for
the four main conventional options used nowadays as an example. Nuclear, peak load, DH CCGT,
and industrial CHP are not included because of their special deployment conditions: nuclear
(Borssele) is always deployed at maximum capacity due to its low fuel cost; peak load units are
deployed only when unforeseen capacity shortages occur, for example, as a result of imbalance; DH
CCGT and industrial CHP units often must run with heat demand being the determining factor for the
units minimum load. Specific heat consumption curves are used for each larger unit in the simulation
model as described in Chapter 8.
11,000

10,500

10,000

Heat Rate [kJ/kWhe]

9,500

Conv Coal

9,000

Conv Gas
HW CCGT

8,500

CCGT

8,000

7,500

7,000

6,500
0

100

200

300

400

500

600

700

Power [MWe]

Figure 4.33: Specific heat consumption curves for various conventional units (Source: SEP, own
Information)
The technology development over the years can be seen in Figure 4.33 for the natural gas fired units
and the change in specific heat consumption. The conventional gas fired units built in the seventys
have a specific heat consumption of approximately 8,780 kJ/kWh (e = 41%). With modification of
conventional units into HW CCGT units in the eighties, the specific heat consumption decreased to
approximately 7,950 kJ/kWh (e = 45%). Later, in the nineties, with the arrival of modern CCGTs, the
specific heat consumption decreased further to approximately 6,790 kJ/kWh (e = 53%). For the
conventional coal-fired units also a change in specific heat consumption can be seen, see Table 4.2,
but the decrease is less sharp compared with the gas-fired units.
Electricity Loss Factor and Co-firing Factor
Usually, CHP units which supply process heat on a continuous basis are large or mid-sized CCGTs with
the option of full condensing operation for large units. Accordingly, these units can in principle be
considered full power plants in the free electricity market. The mid-sized units often have limited cooling
capacity, which explains the need for a minimum heat demand for full-load operation. Most of these units
can be controlled seamlessly. Supplying process heat comes at the expense of part of the generated
electricity, whereby the electricity loss is determined by the pressure level at which the heat (steam) is
supplied. For many processes, such as paper, salt and chlorine production, steam at a pressure level of
around 4 bar is sufficient. This will decrease electricity production by approximately 60 kWh per GJ. The
decrease in electricity production due to heat supply is called the loss factor, and the use of additional fuel
in another power plant to offset this decrease is called the co-firing factor. A loss factor of 60 kWh per GJ
87

Part 2

means that the generation of 1 GJ of heat will result in an electricity loss of around 0.22 GJ (60 kWh x
0.0036 GJ/kWh) which has to be offset elsewhere and will cost 0.4 GJ of fossil fuel (co-firing factor) at
an electrical (CCGT) reference efficiency of 54%.
Most district heat-supplying CHP units are also (larger) CCGTs with the option of full condensing
operation. But because of the (much) lower heat demand in summer there is more scope for on-off
operation as opposed to units which must supply process heat on a continuous basis. Depending on the
temperature and the availability of a district heating coil (a heat exchanger in the outlet part of the HRSG
before the chimney to further cool down the flue gases), electricity production will be decreased by
approximately 0 - 60 kWh for each GJ of supplied heat. This means that the generation of 1 GJ of heat will
result in a co-firing factor between 0 and approximately 0.4 GJ of fossil fuel at an electrical reference
efficiency of 54%. The average annual value of the co-firing factor for a CCGT with adequate heat load is
around 0.3 GJ at a feed temperature of 70 - 120C. The maximum value for modern installations is around
0.35 GJ for every GJ of heat, though this can be much lower at lower temperature levels and in specific
situations. The electricity loss curves caused by heat delivery are used for each larger unit capable to
deliver heat in the simulation model (see Chapter 8).
Combined heat and power applications in the greenhouse and horticultural sector are small-scale and often
involve gas engines. There is no loss factor as the heat is not supplied at the expense of electricity
generation. The first question that comes to mind is how this option performs energetically. At full load a
modern gas engine has an operational electricity efficiency of around 38% and a heat efficiency of
approximately 50% (this can rise to around 57% in the case of low-temperature applications in horticulture).
So, given a fuel input of 100, this installation delivers 38 units of Electricity (E) and 50 units of Heat (H). If
we base the share of fuel attributable to electricity generation on our consistently applied reference norm of
54%, it will be 70.4. So, the share of fuel for Heat is 29.6 (100 70.4) for 50 units of heat, and 1 GJ of heat
produced by the gas engine will cost 0.59 GJ of primary fuel almost double the value of the heat supplied
by a large-scale CCGT as described above. On top of that, the drawback of a gas engine with limited or no
heat supply is its low efficiency in relation to the reference, with the potential risk of transforming the
environmental benefit into an environmental disadvantage. This is why gas engines in, for example,
Denmark, that do not always fully utilize the heat, do not qualify for incentive schemes for heat distribution.
Start-Up Times, Minimum Up and Downtime
To compare the different options, Table 4.14 provides an overview of start-up times (cold start),
minimum up and down times, and controllability (3% of net power range of the units).

Conventional coal
Conventional gas
Hot windbox CCGT
Nuclear3
Peak, Blackstart
IGCC4
CCGT
IND CHP
DH CCGT
Gas engine
Renewables
Waste incinerator

Start-up time (h)


6
5
5.5
NA
<1
24
2
2
2
<1
NA
24

Min Up (h)
24
6
4
NA
1
24
4
4
4
NA
NA
NA

Min Down (h)


24
6
4
NA
1
24
4
6
6
NA
NA
NA

dP/dT (MW/min)
15 - 20
10 - 20
10 - 20
14
1-2
8
4 - 10
3 - 14
1-8
NA
NA
NA

Table 4.14: Some characteristics of the various supply options (Source: Sep, own information)

Nuclear units like Borssele are shut down once a year for a few weeks for the change of a quarter of the nuclear
fuel elements. This period is also used for scheduled maintenance.

The gas turbine of the IGCC unit is first fired with natural gas and later the gas turbine can switch to syngas.
Start-up times on natural gas equals start-up time of the CCGT units but operation on only syngas is depending
on start-up time of the gasification system (typical 24 h).

88

Chapter 4 Energy Supply Chain and Supply Options

The table clearly shows that conventional coal-fired and IGCC units are not suitable for daily on-off
operation, due to the minimum uptime and downtime of 24 hours. As long as the coal price is lower
than the gas price these are real base-load units. In 1999, when the gas price was extremely low,
coal-fired units were usually deployed in control mode, connected to the grid during working days and
operating at minimum load during the night and in weekends. If they could be missed, they were taken
out of operation.
Maintenance
Maintenance-free installations would be perfect but in reality they do not exist. All machines with
moving parts need some maintenance from time to time. No maintenance will irrevocably lead to
unplanned outages a situation which must be prevented. Though maintenance temporarily puts an
installation out of operation, it can be planned and anticipated. Additionally, good maintenance will
reduce the risk of unforeseen outages.
The key aspects of maintenance:
1. Most countries have laws that require periodic inspections of closed containers under
pressure, such as vessels and wide-diameter pipes operating under high pressure and/or
temperatures. In the case of power plants these laws apply to all boilers, main steam pipes,
pressure vessels etc. These inspections have to be carried out by independent professionals
or by company staff with an independent position in the organization and according to a fixed
scheme that observes predetermined intervals.
2. There are no maintenance rules for turbines, but manufacturers provide guidelines.
Steam turbines: all parts should be inspected every 6-8 years and reconditioned if
needed.
Gas turbines: a small to medium inspection is required every 4,000 full-load hours; the
machine should be opened every 12,000-24,000 hours and hot components replaced
by new or reconditioned components.
Gas turbines: compressor fouling leads to decrease in pressure ratio and for example
5% decrease in pressure ratio leads to a decrease in power of approx. 11% and
efficiency of approx. 4.5% (Thomassen International, 1996). With online and offline
(crank soak) washing programs this loss of power and efficiency can be reduced
significant but not completely.
3. In view of the above, coal-fired power plants must undergo scheduled maintenance once
every two years, i.e.:
The steam turbine must be part-opened in accordance with the maintenance scheme.
The insulation must be removed from hot components in accordance with the
inspection plan and submitted to a surveyor (e.g. the Steam Authority) for inspection.
4. As the gas turbine forms the heart of a CCGT, dates and technical support are determined in
consultation with the supplier. Steam Authority inspections are integrated with this scheme.
Combined units have both a large steam turbine and a gas turbine, so the maintenance
schemes are similar to those of coal-fired power plants and CCGTs.
5. Power plants with low variable costs but high operation times are inspected during
maintenance stops in which 24 hours and 7 days a week of working schemes are applied .
Because of the redundant layout of the most critical subparts of these plants maintenance on
these part can be carried out also in full operation.
4.8

Environmental Situation

Table 4.15 lists the average emissions for each option, including the required/possible environmental
measures to reduce them. The scope for process integration, for example co-firing of biomass, steam
from another unit, or gasification, is indicated in Table 4.16 by + or . The need for cooling water, the
site requirements and environmental costs are also briefly addressed.

89

Part 2

CO2
(g/kWhe)

Conventional coal

825

NOX
With /
without
Denox
(mg/kWhe)
600 / 1200

SO2
(mg/kWhe)

NOx
measures

SO2
measures

400

DeNOx,
burners
Burners
DLN, SI, WI,
burners
NA
NA
Nitrogen
DLN
DNL, SI, WI
DNL, SI, WI,
burners
SCR
NA
DeNOx

RIO

Conventional gas
Hot windbox CCGT

490
450

500
450

NA
NA

Nuclear
Peak
IGCC
CCGT
IND CHP
DH CCGT

0
775
780
380
< 380
< 380

NA
180
200
250
300

NA
NA
80
NA
NA
NA

Gas engine
Renewable
Waste incinerator

< 380
NA
-

NA
-

NA
NA
-

NA
NA
NA
NA
Desulfurizer
NA
NA
NA
NA
NA
WetWashing

Table 4.15: Average emissions and required / possible environmental measures (Source:
Environmental Reports, own information)

Conventional coal
Conventional gas,
BFG
Hot windbox CCGT
Nuclear
Peak
IGCC
CCGT
IND CHP
DH CCGT
Gas engine
Renewable
Waste incinerator

Co-firing solid
biomass
++
-

Co-firing liquid
biomass
+
++

Biomass
gasification
+
+

Feeding steam

--NA
----NA
++

++
--NA
--+
NA
-

+
--++
+
++
NA
-

+
-+
+
-NA
+

+
+

Table 4.16: Scope for process integration (Source: Own information)


Notes to Table 4.15: the listed emissions are average values for the options mentioned (with CO2
emissions for CHP options corrected for heat delivery). The possible environmental measures refer to
the most widely used alternatives, some of which were explained when discussing the production
process. For gas turbines the most popular NOx reduction technology is a Dry Low Nox burner (see
Section 6.2.4), but customized burner designs are also applied. The design of conventional coal- and
gas-fired power plants and HW CCGT, in particular, allows co-firing of biomass solid biomass in
coal-fired power plants and liquid biomass in gas-fired power plants. These installations are designed
not only for a primary fuel but also for a secondary and tertiary fuel. This possibility rarely exists in
units based on gas turbines (peak, CCGT and DH CCGT). The gas turbine burners of these units are
designed for one type of fuel, natural gas. The HRSGs are not equipped for co-firing and the steam
production capacity is fine-tuned to the heat input of the gas turbine. Adjustments usually require
substantial investments without any real prospect of good returns or without sufficient guarantees as
regards environmental government schemes. Industrial CHP plants, on the other hand, do have
facilities for burning residual gases. The supplementary firing also increases the steam capacity of the
HRSG, enabling steam from other processes to be (partially) integrated.

90

Chapter 4 Energy Supply Chain and Supply Options

4.9

Security of Supply

The security of electricity supply depends on various factors, one of the most important being installed
capacity versus maximum load. The installed capacity varies in time and is determined by the number
of new development projects and the decommissioning and demolition of older plants. The new
development and decommissioning scenarios are briefly discussed to show that the process of
planning, realizing and commissioning a new development project takes quite some time and that
demolition can lead to re-use of a site for new development.
New Development
The process from planning the construction of a new power plant to its commissioning:
Feasibility, tendering, permits
1. Before coming up with new development plans, in most cases an in-depth feasibility study has
been carried out covering all the financial aspects of the construction and commissioning of
the power plant. In addition to capital expenditure and associated financing, this study
addresses the anticipated electricity proceeds, fuel costs, maintenance costs, other costs, the
impact and reliability of legislation and regulations etc. It takes at least six months to carry out
a thorough feasibility study, also because of higher-level decision-making as to whether or not
the next phase should start.
2. If the decision is positive, the next tendering phase can be launched. The key question in this
phase is whether or not the project is to be subcontracted turnkey or in parts. The latter
scenario leaves overall project management with the commissioning party. Both scenarios
require the compilation of a bid book detailing all the conditions that have to be met by the
subcontractor(s) in the bidding process. Including the assessment of the bids this phase will
also take at least six months.
3. After selecting the subcontractor(s) the actual realization can start. By then, applications will
have been submitted for some permits, others will still need to be requested. The construction
of large plants is subject to MER environmental reporting officially a 7 - month process which
takes longer in practice. The permits are usually granted soon as the MER report has been
approved, though coal-fired plants have to wait longer for them than gas-fired plants.
4. In recent years environmental organizations have lodged official protests with the Council of
State against almost all permit applications, even for gas-fired units. The permit procedure for
the Sloe power plant (gas) has been awaiting a decision for as long as 18 months. If the
Council of State throws out an application, the whole process has to start again from scratch.
The alternative is a provisional permit, which at least allows the work to start. However, civil
proceedings will inevitably follow after a period of months or even years and could still lead to
the withdrawal of the provisional permit.
Construction times5:
1. CCGT plants: 24 months for the standard design, including all the engineering.
2. 600 MWe coal-fired power plants: 42 months.
3. 1,500 MWe nuclear power plants: 6 - 10 years.
The need to drive piles deep into the ground in the Netherlands adds around two months to the
construction time.
Demolition of Power Plants
The demolition of power plants is normally not a problem. The timescale does, however, depend on
whether the entire foundation needs to be removed. Most demolition projects take a year, but it is all
down to the skills of the firm in the long run. Sometimes gas or steam turbines or other leftovers are
re-sold or taken apart instead of scrapped. If this happens, demolition takes a bit longer.

Currently (2008) with high tension on the electric power plant construction market, these construction times have
increased.

91

Part 2

4.10

Summary

The existing technology (applied cycles and main characteristics, applied fuels, production process,
flue gas cleaning, and residual products (if applicable)) of the three main categories conventional
and combined cycle technology, combined heat and power technology and sustainable production
technology of the supply options have been studied in detail in order to get a good starting point for
the reference case for the model of the Dutch electricity supply.

92

Chapter 5 Transport and Distribution of Energy

Chapter 5
5
Transport and Distribution of Energy
5.1

Introduction

This chapter summarizes the present day applied technical configuration for the transport and
distribution grids for electricity, gas and district heat in the Netherlands. The main technological
developments are also summarized in case these developments are of importance and can serve as
direct input for the Dutch electricity supply model. In Section 5.2, the transmission and distribution
grids used for the transport of electricity in the Netherlands are treated followed by a description of the
available interconnector capacity with the surrounding countries and future developments of this
interconnector capacity. This section ends with an analysis of the consequences of a possible large
future contribution of distributed generation in the electricity supply for the energy losses in electricity
grids. In Section 5.3 the transport and distribution grids for the transport of natural gas are treated
followed by a description of the ability to store gas as liquefied natural gas (LNG). This section ends
with a description of the developments for underground storage and the planned extensions in the
interconnector and transport grids. The last Section 5.4 in this chapter is a brief description of heat
distribution grids, their setup and their place in the energy chain. This section ends with the future
developments for this grids in case these developments are of importance for the Dutch electricity
supply model. District cooling has recently been introduced in the Netherlands and is a promising
energy-saving/environmentally-friendly technology for the future which is why this technology is
discussed as well but to a limited extent only.
5.2

Electricity Transmission and Distribution Grids

5.2.1

Introduction

The primary function of the electricity infrastructure is to link the demand and supply of electricity as
efficiently as possible.

Figure 5.1: Supply and transmission of electricity at the Amer power plant in Geertruidenberg (Source:
Essent)

93

Part 2

In its most elementary form this should result in an energy supply that directly links demand and
supply. For example, large electricity consumers that establish sites near a hydroelectric dam a coal
mine or a source of natural gas.
This is how energy was first supplied in many places, but the current situation is very different for the
following reasons:
1. Electricity was not available until the end of the nineteenth century. At that time, large parts of
the world were already settled.
2. Electricity was so cheap that companies did not even consider it when deciding where to
establish sites. Moreover, and in particular in the developed areas, electricity became
indispensable to humans. Electricity must be available wherever humans decide to settle for
whatever reason.
3. Fuel and electricity can be transported over large distances. However, transporting electricity
over large distances (>250 km) is often the most expensive option (Van Eck et al, 2004).
4. There are more and more ways of generating electricity and the large-scale storage of
electricity can be a possible future option (see Section 6.4.5).
For the future situation of electricity transport and distribution three main tendencies are possible:
1. Focus on large-scale conversion and supply of electricity and electrical infrastructure systems.
2. Focus on small-scale and local solutions for supply of electricity, also known as "Distributed
Generation (DG)".
3. A combination of both.
5.2.2

Transporting and Distributing of Electricity

Electricity generated at central production facilities is, after transformation to the required voltage,
transported at high (150, 100 and 50 kV) or ultra-high voltages (380, 220 kV) over large distances
towards the several switch- and/transformation points. After transformation (in general to 10 kV and
0,4 kV) the electricity is consumed (see Figure 5.2). The ultra-high voltage grid is also used for import
and export of electricity. It is therefore necessary to transport electricity from the higher levels of the
grid to the lower levels of the grid. This is done via the transmission and distribution grids.

Figure 5.2: Distribution of consumption and supply on the grid levels (Source: TenneT, 2002)

94

Chapter 5 Transport and Distribution of Energy

5.2.3

Structure of the Transmission Grids

The transmission grids and interconnectors are displayed in Figure 5.3.

Figure 5.3: Dutch transmission grids and interconnectors (Source: TenneT)


The 380 kV and 220 kV transmission grids usually transport large amounts of electricity over long
distances from a limited number of generators, or import/export electricity to a limited number of
switch- and/transformation points (see Figure 5.3). Every 25 to 50 km there is a transformation point to
a sub-transmission grid with a lower voltage. In the Netherlands, this is the 150 kV, and 110 kV highvoltage grid. At several places the 150 kV grid is coupled with a transformation point to a 50 kV grid. In
these 150, 110 and 50 kV sub transmission grids, every 10 to 15 km there is a transformation point to
medium voltage 20- or 10 kV and next in this grid every 0.5 km to 1 km a transformation point to low
voltage (TenneT, 2002). The transmission grids fall within the voltage level category of 50 kV and
higher, and consist of overhead lines with masts, switch- and/transformation points and underground
cables.
Generators can feed in at all voltage levels of the transmission grid, but the large production facilities
as Borssele, Eems, Geertruidenberg, Maasbracht, Maasvlakte, are coupled directly with an ultra high
voltage switch- and/transformation point and the other large and medium production facilities
commonly feed in directly at high voltage grids.
5.2.4

Structure of the Distribution Grids

All of the distribution grids have a voltage level of 20 kV and lower (10 kV and 0.4 kV). In contrast to
the transmission grid, the distribution grid consists of a fine-meshed underground network of cables
with numerous branches, circular lines, substations, etc. The construction and operation of these grids
largely depend on the supply of power from higher voltage grids and the connection of customers.
There are a number of areas in which so much capacity is supplied at this voltage level (in particular
by gas engines and windmills) that the relevant distribution grids can export the overcapacity. This is,
however, paired with issues, such as grid stability, voltage management, short-circuit capacity, and
security of supply but is outside the scope of this research.

95

Part 2

On the customer side, there is a lot of unused potential for demand-side management (Van Eck,
2007). Should this potential be used, and providing the focus is on distributed generation, for supply of
electricity and feed in on the distribution grids, it will have an impact on the way these grids are
designed and operated. The influence of DG on the energy losses in the grids is briefly analyses in
Section 5.2.6 as input for the Dutch electricity supply model which includes a scenario with a large
amount of distributed generation. The consequences of DG for development of the the grids are
however outside the scope of this research.
5.2.5

Cross-Border Connections

Connections with Germany and Belgium


Present-day (2008) there are five 380 kV alternating current (AC) cross-border connections, the socalled interconnectors (see Figure 5.3). Three with Germany and two with Belgium. Each connection
consists of two circuits that can be used separately, see Table 5.1.There are a few other cross-border
connections at lower voltage levels. Because the possibilities to transport electricity across these
connections are marginal, they are not taken into account.
The Hengelo-Gronau connection can directly control the transmission flows by means of a phase
shifter on the German side of the connection. In 2003, similar phase shifters went into operation in the
Meeden-Diele/Conneforde connection.
Current Capacity
At the end of 2003, the installed connection capacity was distributed across the interconnectors as
follows:
Connection

Type

Meeden-Diele/Conneforde

2 circuits and 2 phase


shifters
2 circuits and 1 phase
shifter
2 circuits

Hengelo-Gronau
MaasbrachtSiersdorf/Rommerkirchen
MaasbrachtMeerhout/Gramme
Geertruidenberg-Zandvliet
Borssele-Zandvliet

2 circuits
1 circuit
1 circuit via a
380/150kV transformer

Thermal capacity
[MVA]
2 x 1,370

Operator

2 x 710

TenneT - E.on Netz

2 x 1,710

TenneT - RWE Netz

1 x 1,350
1 x 1,420
1 x 1,645
1 x 450

TenneT - ELIA

TenneT - E.on Netz

TenneT - ELIA
TenneT - ELIA

Table 5.1: Connection capacity across interconnectors and operators (Source: TenneT, 2002)
The installed connection capacity totals 12,445 MVA.
To determine the securely available cross-border transmission capacity for the electricity market,
TenneT takes the following criteria into account:

96

The single failure criterion (see Netcode). This so-called (n-1) criterion means that a cross-border
connection must be able to fail at any time in the Netherlands without creating transmission
problems.
The upper limit of the connections' loads, which is agreed with the foreign grid operators.
The outcome of the load-flow calculations to determine the power flows on the cross-border
connections.
The minimization of reactive power exchange as agreed within the UCTE framework.

Chapter 5 Transport and Distribution of Energy

With the installation of the two phase shifters in Meeden, the available capacity was expected to
increase (TenneT, 2002) as follows:

To 3,900 MW on February 1, 2003, after the second phase shifter went into operation.
To 4,600 MW on April 1, 2003, after the successful completion of the trial period with the two
phase shifters.
To 5,000 MW starting mid-2003, after voltage and reactive power management measures had
been implemented.

The present-day Dutch interconnection is, in other words, already capable of importing a total of 5,000
MW. After deducting the reserved capacity of 300 MWe for the TSO (required as a result of
agreements on assistance power within the UCTE framework (TenneT, 2004)), 4,700 MWe should
be available for the electricity market. The following factors, however, impact the expected securely
available cross-border transmission capacity for electricity:

Planned maintenance and failures in the Dutch 380/220kV grid including the cross-border
connections.
Failures with consequential loss in the high-voltage grids of the TSOs in the surrounding countries.
Unpredicted parallel power flows that are caused by electricity market transactions in other
countries and that increase the load of the Dutch 380/220kV grid and cross-border connections.
Grid limitations in another country that are caused by highly variable current patterns in Northwest
Europe due to the inflow of power from the numerous wind turbines in North Germany.

This is why the additional anticipated import capacity that is supposed to be available for the electricity
market is not reached. At the end of 2004, the capacity released to the electricity market was still only
3,350 MWe, which, the operational situation permitting, was increased to 3,600 MW and respectively
3,850 MWe. The main cause is the highly variable international current patterns, which are caused in
particular by the strongly fluctuating supply of wind energy in North Germany. Figure 5.4 shows that
50% of the German onshore wind power6 supplies E.on's North German grid.

Figure 5.4: E.on's North German grid (Source: E.on Netz, 2003)

At the end of 2003, the total installed onshore wind capacity in Germany was approximately 15,000 MWe. This
increased to 22,247 MWe at the end of 2007 (German WindEnergy Association (BWE), 2008).

97

Part 2

In case large amounts of electricity are generated in North Germany, whilst the demand of electricity is
mainly in the middle and the southern part of Germany, in combination with the restriction in
transmission grids from North to Southern Germany, this leads to variable current patterns resulting in
the so-called transit flows: the physical movement of electricity through the Netherlands, Belgium,
France, and back to Southern Germany. To control these transmission situations, TenneT and the
TSOs in the surrounding countries are implementing a number of measures, one of which is the
installation of new phase shifters, increase transmission capacities7 and, if required, the limitation of
the import capacity allocated to the Dutch electricity market. According to TenneT (2004), the recently
installed phase shifters are serving their purpose because power flows can be controlled. It is stated
that the current level of import would have had to be decreased considerably, several hundreds of
megawatts, if the phase shifters had not been installed. Upgrading of the internal grids in Germany
has a positive effect on the available cross-border capacity.
NorNed Cable
General
In this project, which was developed by TenneT and Statnett, a High-Voltage Direct Current (HVDC)
electricity cable is constructed between Norway and the Netherlands with a transmission capacity of
about 600 - 700 MWe. The total project cost are estimated on 600 million (TenneT, 2007) and the
project was started in 2005 and has come into service on May 6, 2008 (TenneT, 2008).
System Structure
All of the existing cross-border connections described previously are AC connections. The distance
between Norway and the Netherlands however is substantial (the route from Eemshaven to Freda
(Norway) has a length of approximately 580 km), meaning that the distance for a submarine cable
connection based on AC would be too large and a HDVC connection is chosen. AC can not take
power over that distance, ACs technical limit is 100 km before losses begin to grow to unacceptable
levels (Skog, 2006). One of the benefits of HDVC is related to its effect on grid reliability. HDVC lines
cannot be overloaded and power flow can be controlled for grid stability. Another benefit is the greater
power flow control in respect to the direction and capacity which can be afforded to the electricity
market and system operators (EPRI, 2006). For the Norned cable a bipolar direct-current connection
with two cables (in case of failure of one cable, 50% capacity is retained) insulated for the full voltage
has been chosen. The selected transmission voltage is +/- 450kV DC.
The system is structured as follows (Norway - Netherlands):
Norwegian AC grid in Feda (municipality of Kvinesdal).
Connection between AC grid and converter station.
Converter station.
Submarine cable.
Converter station in Eemshaven.
Connection between converter station and AC grid.
TenneT 220/380kV transformer station Eemshaven.
Transmission configuration
The main circuit configuration consists of a 12 pulse converter 450 kV with the midpoint earthed and
two cables. The transmission voltage is effectively 900 kV giving fairly low cable current and losses.
The DC link is designed to operate continuously at 700 MW when all converter cooling equipment is in
operation. The configuration is shown in Figure 5.5.

With the installation of the new 400 kV Avelgem (BE) Mastaing (FR) interconnection line and the new phase
shifter in Monceau the maximum transmission capacity values from France to Belgium have increased form 2,250
MW to 3,200 MWe. To secure the Belgian grid as well as the Central Western Area, 3 phase shifters will be
installed in the spring of 2008. One of them will be installed at Zandvliet (BE), the two other at Van Eyck (BE)
(UCTE, 2008).

98

Chapter 5 Transport and Distribution of Energy

Figure 5.5: NorNed main circuit configuration (Source: Skog, 2006)


Converter stations
In Feda the AC side equipment is connected to the 300 kV substation and placed outdoors, see Figure
5.6. In the Eemshaven the AC side equipment is connected to the 380 kV station by cables and the
AC filters are located indoors. In Eemshaven also the DC side equipment is installed indoors because
of the continuous slat deposit. Converter stations are necessary because AC has to be converted to
DC and vice versa. For the NorNed project, the converter transformers are of single phase, three
winding type. Six double-valves in the valve hall are arranged to provide the 450 kV twelve pulse
convetre. The twelve-pulse convetre has a voltage rating of 900 kV while the DC voltage to ground is
450 kV. Each single valve has 120 thyristors including three redundant devices (Skog, 2006). Filter
banks consisting of coils and capacitors are installed to prevent "polluting" the 220 and 380kV grid. For
an extensive description of the converter station please refer to the applied literature (Skog, 2006,
Bahrman, 2007).

Figure 5.6: Feda converter station, Norway (Source: ABB)

99

Part 2

Cable Technology
Use is made of the universally applied conventional cable technology based on paper-insulated,
mass-impregnated cables with copper conductors, a lead casing, and a double steel-wire
reinforcement.
Losses
At 600 MW, the connection has a 3.7% loss (Skog, 2006), and at 700 MW, a 5.5% loss (TenneT,
2004).
Life Span, Operation and Maintenance Costs
The expecting life span is set to 40 years without any major reinvestments having to be made. The
annual operation and maintenance costs, including insurances, for the NorNed connection are
estimated at 4 million (TenneT, 2004).
Consequences for the current Cross-border Transmission Capacity
TenneT calculated the load-flow effects of the NorNed cable for feeding a 600 MW cable (TenneT,
2004). The conclusion is that a 600/700 MW cable can be fully utilized without decreasing the
available cross-border transmission capacity. The Dutch grid, however, will have to be reinforced by a
second 380/220kV coupling transformer of 750 MVA in the Eemshaven station in order to transport
the capacity safely under all circumstances.
The Future
The connection can be expanded to 1,200 MW. However, Statnett (2004) recommends to first
implement the 600 MW connection, which has an overload capacity of 100 MW. Providing it is
economically attractive, the expansion can take place at a later stage by laying a new cable in parallel
to the existing cable connection.
Cable to England
BritNed, a joint venture between TenneT and National Grid, have been developing a project for a
submarine HVDC electricity cable between the United Kingdom and the Netherlands and announced
in May 2007 that this link is to be constructed and operational by late 2010. This interconnector will
have a capacity of 1,000 MW and will be 260 km long and will run beneath the North Sea between the
Isle of Grain in Kent to Maasvlakte, near Rotterdam. The total project cost are estimated on 600
million (TenneT / National Grid, 2007).
5.2.6

Near Future Developments

Transmission grids
In addition to the described NorNed cable and the cable to the United Kingdom, the construction of
new coal- and gas-fired units, offshore wind farms, and increased export and import will have the
biggest impact on the transmission grid. These developments should give rise to the creation of new
connections and the replacement of existing connections with a higher capacity (TenneT, 2002). The
following projects are particularly important (TenneT, 2006):
The Randstad380 project, which involves the extension of the grid from the Maasvlakte to
Beverwijk via Bleiswijk.
The Diemen-Oostzaan-Beverwijk project, which involves the extension of the grid and
construction of the 380kV stations Oostzaan and Beverwijk to strengthen the connection to the
Randstad .
The reactive-power compensation-measures project which involves condenser banks and
coils that enable the consequences of the increasing power transports over the national highvoltage grid to be compensated.
The phased expansion of the 380kV station, Borssele, in order to facilitate developments
within an acceptable time frame, such as the connection of generation units and the
improvement of the import capacity with Belgium.

100

Chapter 5 Transport and Distribution of Energy

Distributed Generation
Electricity produced in the OECD countries is generated mainly in large electricity generating units
which produce and transmit electricity through ultra and high-voltage transmission systems then, at
reduced voltage, transmit through local distribution systems to consumers. Distributed generation
(DG) is a generating unit serving a customer on-site or providing support to a distribution network,
connected to the grid at distribution level voltages, see Figure 5.7. The technologies generally include
gas engines, small (and micro) gas turbines, fuel cells, and photovoltaic systems. It generally excludes
wind power, since that is mostly produced on wind farms rather than for on site power requirements
(IEA, 2002).

Figure 5.7: Distributed Generation in an electricity network (Source: IEA, 2002)


DG plants already represent a small share of the electricity supply in the OECD countries and play a
key role for applications in which reliability is crucial (as a source of emergency capacity), as an
alternative to expansion of a local network and as the combination of heat and power generation
(CHP). The CHP generation has the potential to alter fundamentally the structure and organization of
our electric power system (IEA, 2002) and is already applied on a large scale in the Netherlands,
Finland and Denmark. DG, depending on location may offer additional value to the grid :
Deferral of upgrade to the transmission system.
Deferral of upgrade to the distribution system.
Reduction of losses in the distribution system.
Provision of network support or ancillary services.
On site production avoids transmission and distribution costs, which other wise amount to about 30%
of the cost of delivered electricity (IEA, 2002). According to Dondi (2002) cost savings of 10 - 15% can
be achieved. However these results are only correct when the distributed generation units are standalone units and dont appeal to the grid. If not they are jointly responsible for the distribution grid and
its losses (Pepermans et al, 2005; Pauwels et al, 2000).
The average of grid losses in the OECD countries amounts 6.8% (IEA, 2002). The question if and how
much grid losses are saved by the use of DG is dependent of the location of the DG, the amount of
DG applied but also of the applied net configuration of the location. Also the production patterns of the
DG plant are of importance in relation with the consumption pattern. If electricity is produced at time
with no consumption, often the case for photovoltaic systems and wind energy (wind is however not
DG) or more electricity is produced then required by the local consumer, the electricity is direct
delivered via the distribution grids to the transmission grid and transported to places where the
electricity consumption takes place. The benefit of the DG plant close to the consumer and with a

101

Part 2

minimum of grid losses does not apply in that case because electricity has to follow a long route to the
consumers (De Jong, 2003). Figure 5.8 shows the relation between the share of DG in the total
electricity production versus the grid losses for the countries of the European Union.
10.0
Portugal

Greece

9.0

9.0

8.8

8.6 Spain

Ireland

8.0

8.1

7.2 Sw eden
6.9 Italie

7.0
Distribution Losses [%]

UK

8.0

6.0

France

5.6

5.0
Belgium

5.5 Germany
4.7

Denmark

4.7

Netherlands

Austria

4.4

4.3

4.0

Finland

3.4
3.0

2.0
Luxembourg

1.2

1.0

0.0
0.0

10.0

20.0

30.0

40.0

50.0

60.0

70.0

80.0

Share CHP of total electricity production [%]

Figure 5.8: OECD Countries Share DG versus Distribution Losses (Source: IEA, Eurostat)
As can be seen in Figure 5.8 the countries with a large share of DG have the lowest % of grid losses
but above a share of 20% no clear decline is noticed by increase of the share DG also by the absence
of more data. The grid losses in the countries with a share of DG bellow 20% have a range between
4.7 and 9.0%. For this countries the applied net configurations, extensiveness of the country (Italy,
Sweden) are of importance. For the Netherlands distribution losses of 4.3% apply for a share of DG of
approximately 29.5% and it is expected that a further increase of the share of DG in total electricity
production leads to a further decrease in % grid losses.
Demand Side Management and Economics of Infrastructures
These developments get more attention because the share of distributed generation increases,
particularly the intermittent type such as wind, solar, small hydro and combined heat and power (small
and micro-CHP). Due to the fact that intermittent types of electricity generation are difficult to predict,
electrical networks - both local and distribution/transmission - are turning to integrated distributed
energy resource. By combining distributed generation with energy storage and demand response,
countries can decrease problems caused by distributed generation and increase the value of
intermittent energy in the market. The IEA Demand Side Management Programme is Promoting
Energy Efficiency and Demand-Side Management for global sustainable development and for
business opportunities and is one of more than 40 co-operative energy technology programmes
sponsored by the International Energy Agency (IEA).
Economics of Infrastructures deals with the design of efficient and effective governance structures for
infrastructure industries (like ICT, energy, water, transport and logistics) at the interface of economics
and technology.
Research for these developments continues but this is beyond the scope of this thesis. For more
detailed information about these developments, please refer to the literature and relevant websites of
IEA (dsm.iea.org) and University of Delft (www.ei.tbm.tudelft.nl).

102

Chapter 5 Transport and Distribution of Energy

5.3

Gas Transport and Distribution Grids

5.3.1

Introduction

In 1948, natural gas was found near Coevorden in the Dutch province of Drenthe. In 1951, the
residents of Coevorden were the first to use natural gas in the Netherlands. In the late 1950s, natural
gas was discovered in Slochteren. The gas field was so big that it was considered to be one of the
largest fields in the world and to be of great importance for the Dutch economy. With his 1962 Nota
inzake het aardgas ("Explanatory Memorandum on Natural Gas") the Minister of Economic Affairs,
Jan de Pous, created the framework for the development of the natural gas sector in the Netherlands
(see also the Introduction). Other fields were found later, both onshore and offshore in the Dutch
sector of the North Sea (see Figure 5.9). Among other things, the "small field policy" enables these
fields to be developed and exploited.

Figure 5.9: Gas fields and gas grid infrastructure in the Netherlands (Source: Correlj et al, 2003)
( Theo Barten, 's-Hertogenbosch)
By the end of 1963, the first Groningen gas was supplied to consumers, and by 1968 all of the
mainland municipalities were connected to the natural gas grid (Correlj et al, 2003). To get the gas to
the consumers, an extensive gas transport system was built as of the 1960s (see Figure 5.10).
5.3.2

Transport and Distribution of Natural Gas

The primary function of the gas infrastructure is to get the gas to the consumers as efficiently as
possible. Consumers are bulk consumers (industries, electricity generation units), private consumers,
and international consumers (export). Natural gas is extracted under high pressure, but most
consumers use low pressure gas. It is therefore necessary to reduce the pressure of the gas so the

103

Part 2

consumers can use it. Today's extensive gas transport system, which consists of a main transport
system (MTS), a regional transport system (RTS), different types of stations (operated by Gas
Transport Services (GTS), the national grid operator for natural gas established in 2004), and the
regional distribution grids (operated by the energy companies), is set up to do this.
The gas transport system is made up of the following components (Source: N.V. Nederlandse
Gasunie, 2005):
Main transport system
Regional transport system
Metering and pressure-regulating stations (77)
Export stations (10)
Compressor stations (9)
Blending stations (11)
Custody transfer stations (about 1100)
LNG storage facilitie (1)
Calibration stations (2)
Nitrogen plant (1)
Regional distribution grids.
Main Transport System
The main transport system (see Figure 5.10) is supplied from the Groningen gas field (maximum
working pressure between 67 bar (g) and 81 bar (g)), the various blending stations that process and
supply the gas produced offshore and onshore (see Gas blending stations), and storage facilities for
natural gas (see Storage natural gas). The function of the main transport system is to:
Deliver natural gas to metering and pressure-regulating stations from where it is supplied to
the regional transport system.
Deliver natural gas to custody transfer stations near bulk consumers from where it is supplied
to industries and power stations that are directly connected to the main transport system.
Deliver natural gas to export stations from where it is supplied to consumers abroad.
Figure 5.10 shows that the transport system consists of a transport system for Groningen gas (G-gas
transport system) as well as a separate transport system for high-calorific gas (H-gas transport
system)8. This is due to the active development policy for small gas fields, exploited by gas producers
and Gasunie buying the produced gas (small field policy) so that the Groningen field can be used as
a strategic reserve and flexible gas source for as long as possible. In the 1960s, there was only one
quality of gas that of the Groningen gas field. This changed in the 1970s because gas was being
bought from onshore fields other than the Groningen field and from off shore fields. Compared to the
Groningen gas (containing approximately 14% nitrogen), most of the bought gas has a very low level
of nitrogen (approximately 1%), meaning that they have a high-calorific value. The bought gas is also
not immediately usable as distribution gas because of a different quality factor (Wobbe index). In order
to deliver this gas to the consumers, a number of measures were taken:
A separate H-gas transport system was built for bulk consumers, such as industries and
power stations, and for the export of gas. Existing burners were converted or new burners built
that could burn the high-calorific gas. These measures enable costumers to buy and use highcalorific natural gas.
The composition of H-gas was modified by adding nitrogen to match the quality of Groningen
gas (the other way around is also possible). This is done in the blending stations.

The Groningen gas (G-gas) has a Higher Heating Value (HHV) of 35.2 36.2 MJ/m3 and a Lower Heating Value
(LHV) of 31.7 - 32.6 MJ/m3. The high-calorific gas (H-gas) has a HHV of 39.7 - 44.0 MJ/m3 (GTS, 2008) and a
LHV of (approximately) 35.7 39.6 MJ/m3 . The difference between HHV and LHV is the condensation heat of
water vapor in the combustible products which is included in the HHV.

104

Chapter 5 Transport and Distribution of Energy

Figure 5.10: Schematic overview of the main transport system in the Netherlands at year end 20069
(Source: Gas Transport Services)
Regional Transport System
This transport system (finely meshed structure) is supplied from the main transport system via the
metering and pressure-regulating stations and has the following functions:
Deliver natural gas to the custody transfer stations of energy companies from where the gas is
delivered to bulk consumers and private consumers via the energy companies' regional
distribution grids.
9

Not on the schematic overview of the main transport system in the Netherlands is the ZEBRA gas pipeline from
Zelzate to Ossendrecht.

105

Part 2

Deliver natural gas to the custody transfer stations near bulk consumers from where it is
supplied to industries and electricity generation units that are not directly connected to the
main transport system.
Deliver natural gas to export stations (small portion) from where it is supplied to consumers
abroad.

The length of the entire main and regional gas transport system is approximately 11,600 km.
Metering and Pressure-Regulating Stations
The metering and pressure-regulating stations divide the main transport system and the regional
transport system and have the following functions:
Transfer gas from the main transport system to the regional transport system at a reduced
pressure of 40 bar, the regional transport system's maximum allowable working pressure.
Odorize the odorless natural gas to give it a characteristic smell so that it can be immediately
detected and prevent hazardous situations.
Meter the delivered amounts of gas as a means of controlling and monitoring the transport
system.
Export Stations
Export gas is primarily supplied from the main transport system and delivered through export stations.
The gas delivered to consumers abroad is metered in these stations in order to provide them with a
financial statement of their transactions. Several export stations are also equipped to import gas.
Compressor Stations
To prevent the pressure in the system dropping too much as a result of frictional resistance,
compressor stations are built in several locations along the main transport system. The power of these
compressor stations, which are often centrifugal and powered by gas turbines, varies from 20 MW to
approximately 175 MW, whereby the capacity required depends on the throughput.
Blending Stations
Blending stations are built in several locations to bring H-gas to the G-gas quality level in a controlled
manner, and vice versa.
Custody transfer stations
Natural gas is delivered to bulk consumers, such as industries, electricity generation units, and energy
companies from the main transport system and the regional transport system via custody transfer
stations, where the pressure of the gas is reduced to the pressure requested by the consumer
(frequently 8 bar). To prevent the gas dropping to undesirably low temperatures after its pressure has
been reduced, the custody transfer station is equipped with a heating system. The gas delivered to the
consumers is metered in order to provide them with a financial statement of their transactions.
Regional Distribution Grids
Gas is distributed to private consumers and small and medium-sized commercial consumers through
the regional distribution grids, which are supplied by the custody transfer stations. The real bulk
consumers are usually directly connected to a custody transfer stations.
5.3.3

Storing Natural Gas

The Groningen gas field's main purpose is to supplement the gas produced onshore and offshore to
ensure that the total demand for natural gas is met during the coldest days. The Groningen field is
also used to balance the supply on other days because the onshore and offshore fields produce a
more or less constant volume (small field policy) but do not have a real swing capacity due to their
small size. This balancing role is a so-called swing role. As a result of producing natural gas since the
1960s, the initial volume of the field has since been halved. This has reduced the field's pressure,
which in turn has continuously reduced the field's maximum supply capacity. Because of this, Gasunie
can no longer guarantee that enough gas will be available everywhere on the coldest days of the year
(NAM and Gasunie, 2001). For this reason, gas storage facilities play an important role in
106

Chapter 5 Transport and Distribution of Energy

guaranteeing the security of supply and delivery because they can supply if the capacity from the
existing production fields cannot fulfill the demand. Natural gas can be stored in LNG terminals or
underground in, for example, depleted gas fields, salt caverns, or aquifers.
LNG Installation
At the Maasvlakte, Gasunie uses an installation that was built in 1977 to liquefy and store (highcalorific) natural gas that is supplied via the main transport system (see Figure 5.11).

Figure 5.11: LNG terminal Maasvlakte (Source: Correlj et al, 2003)


LNG stands for Liquefied Natural Gas. Gas is liquefied by drying it and removing the undesired
pollutants. The next step consists of cooling the gas to -69C whereby the remaining hydrocarbons
condensate and are removed. Further cooling to -100C produces a mixture of methane, ethane, and
nitrogen, whereby the nitrogen is then separated. The remaining gas is further cooled to -169C
thereby liquefying the gas, which can now be stored in the atmospheric tank.
Gasunie's terminal in the Maasvlakte can store approximately 75 million m of liquid gas, which,
compared to the volume of the gas in its original state, represents a reduction in volume by a factor
600. The maximum release capacity is 1.3 million m per hour.
The separated nitrogen is liquefied by cooling it to -196C and stored in a tank. When natural gas is
supplied from the LNG terminal, the nitrogen is vaporized and mixed with the vaporized natural gas
until the right ratio is reached. Because the stored natural gas has a low nitrogen level and the
supplied natural gas must meet the Groningen quality (approximately 14% nitrogen), a considerable
amount of nitrogen must be bought.

107

Part 2

Underground Storage
Additional production capacity can also be created through underground storage. Gasunie currently
has three underground storage facilities in Alkmaar, Grijpskerk, and Norg. Table 5.2 describes the
functionality of these fields. A few gas storage installations that are representative for Germany are
included for illustration purposes.
Location

Cushion gas

Work volume

Alkmaar
Grijpskerk
Norg
Dtlingen
Rehden

[billion m 3]
3.1
11.6
25.0
2.4
2.8

[billion m 3]
0.5
1.5
3.0
2.0
4.2

Cushion gas
Production Operation time
Injection Operation time
:
capacity
production
capacity
injection
3
3
Work volume [million m /day]
[days] [million m /day]
[days]
6.2
36
14
4.5
111
7.7
55
30
12
125
8.3
55
70
24
125
1.2
20
100
9.6
208
0.7
58
72
34
124

Table 5.2: Functionality of Gasunie's underground storage facilities (Source: NAM en Gasunie, 2001)
From the data in the table, we can see that it takes at least 111 days to inject 500 million m of gas
into the Alkmaar storage facility and less than 14 days to release it. Supporting the Groningen field
(security of supply), the facility in Norg is particularly suitable for the supply of additional production
capacity during the winter. In the summer because of the size of the injection process Norg can be
easily brought to the required level through the collection of produced gas from Dutch fields (small
field policy).
5.3.4

Near Future Developments

To ensure that the security of supply of natural gas is maintained and the interests of the Netherlands
as gas producing and transit country (interconnection function) are secured in the future, Gasunie10
signed in recent years a number of large import and export contracts with countries such as the United
Kingdom and Russia. Alliances have also been established and innovative studies and projects
started. In short (sources, among others, Gasunie Transportservices, Energyvalley):
This has resulted, among other things, in the decision to lay a pipeline to the United Kingdom,
the Balgzand Bacton Line (BBL), which required an investment of approximately 500 million
and has come into operation at the end of 2006.
Gasunie and a number of parties are discussing a direct connection between the Russian gas
fields and the Netherlands, the "Baltic Pipeline".
EuroHub (Oude Statenzijl) entered into a partnership with HubCo (Bunde/Emden) thereby
strengthening the European position of this international natural gas hub.
Gasunie and Nuon are building a gas storage facility near Zuidwending in Groningen. The gas
will be stored in underground caverns with a depth between 1,000 and 1,500 m and with a
storage capacity of approximately 180 million cubic meters. The total invesetment cost of the
whole project is estimated at approximately 350 million. It is expected that the gas storage
becomes operational to its full extent at the end of 2009 (Nuon, 2006).
The high-pressure gas grid is being strengthened, i.e. expanded to accommodate larger
transit flows from North to South.
Increased requirements for natural gas will be satisfied also by LNG suppliers. Four LNG
terminals are already on the drawing board for the Netherlands. Once final decisions are being
made, investment in the transmission network will also be necessary.
Gasunie is working on the development of a third-generation high-efficiency boiler, the micro
CHP which generates both heat and power.
Feasibility studies are being carried out to investigate the production possibilities of New
Gas/Green Gas based on biomass, as well as the possibility of adding hydrogen to the
existing natural gas grid.
10

As of July 1, 2005 the Gasunie is unbundled in two separate companies: the commercial company Gasunie
Trade and Supply (renamed as of September 1, 2006 to GasTerra) and (regulated) network company
Nederlandse Gasunie. Gas Transport Services (GTS) is a wholly owned subsidiary of Nederlandse Gasunie.

108

Chapter 5 Transport and Distribution of Energy

5.4

Heat Transport and Distribution Grids

5.4.1

Introduction

A characteristic feature of heat distribution is that there are many consumers, both private and bulk
consumers, and usually only one or very few heat generating plants. Heat distribution is always a local
function that is not physically connected to other grids. In contrast to the electricity and gas grids,
interruptions never affect other areas. This part of the system is discussed to a limited extent only.
5.4.2

Transport and Distribution of District Heat

Within the framework of this research, heat distribution projects refer to projects that have a collective
energy-saving/environmentally-friendly heat source, the heat of which is delivered to several customers
through a heat transport grid (also known as primary grid), a heat transmission station and a distribution
grid (also known as secondary grid) (see Figure 5.12). Hot water is transported and distributed by means
of a dual pipe system in which one pipe is used to supply the hot water and the other is used to return
the cooled down water. It is a closed system that only needs to be supplemented by the water that
leaks. The transport and distribution grid are normally physically separated by district-based heat
transmission stations (also known as control chambers) that have heat exchangers and, depending on the
project, provisions for the delivery of hot tap water.

Figure 5.12: Schematic diagram of a district heating and district cooling system (Source: District Energy)
There is a direct relationship between the temperature and the pressure conditions in the heat distribution
grids and the operating conditions of the heat generator. In particular the temperature level of the
generated heat and return condensate have an impact on the heat source's energy efficiency. A low return
temperature from the distribution grid usually improves the unit's energy efficiency because it uses the lowvalue heat from the steam turbine(s) better. The same applies to a lower supply temperature. This avoids
having to use high-quality steam to produce a high supply temperature so that less electricity is lost.

109

Part 2

Most of the new and recently built distribution grids in the Netherlands are 70/40C grids, which amounts to
a supply temperature at the consumers of 70C and a return temperature of 40C. Hot water must be
supplied to homes with individual hot water tap heat exchangers at a temperature of 70C the whole year
to ensure a minimum tap water temperature of 60C and rule out any risk of Legionella pneumophila. In
districts with a central hot water tap heat exchangers, hot tap water is generated in the heat transmission
stations and transported in separate pipes to the consumers. In this case the supply temperature in the
distribution grid can be reduced bellow 70C during summer conditions. A secondary return temperature of
40C can be achieved in both systems and a return temperature between 45 - 50C to the heat source is
possible.
The choice of the supply temperature from the heat source is determined, on the one hand, by the demand
for heat, and on the other, by the dynamic aspects and limitations that go hand in hand with the transport of
water. The capacity of a heat transport pipe, which is determined by the pipe's inside diameter, is best used
when the temperature difference between the supply and the return is as high as possible so that the hot
water mass flow through the pipe is as low as possible. The physical speed limits in the grid the speed
may often not exceed 3 m/s enable the grid to be dimensioned.
On winter days with a high heat demand, the heat source usually supplies heat at a temperature of 120C.
At a return temperature of 45 - 50C, a lot of capacity is available in the transport grid. In the summer, when
the heat load is low, the supply temperature is set to a lower temperature, namely 80C, to get enough flow
through the transport grid due to a lower temperature difference and to ensure a minimum tap water
temperature of 60C at the consumers places.
The yearly heat losses as % of the total produced heat in DH systems varies depending on the scale of the
DH system, the age of the DH system and if individual or central tap water system is applied.
A large DH system for example supplied with industrial-, waste incinerator- or large scale CHP heat has a
yearly heat loss in the range of 25 - 35%, a smaller DH system for example supplied with small scale CHP
has a yearly heat loss in the range of 5 - 15%. The older and larger DH systems, for example the existing
DH system in Rotterdam and Utrecht still have consumers who have an older 90/70C system instead of
70/40C systems, in this case heat losses in the DH system are at the upper site of the 20 - 35% range.
Recently build DH systems are on the lower site of the 25 - 35% range. Existing systems with central tap
water have less energy losses in the distribution network during summer conditions, and yearly heat losses
are in the middle of the 25 - 35% range.
Because of the extreme heat peak (approximately 70% of the maximum load consists of a heat peak (see
Figure 4.15 of Chapter 4)), the transport grid can be designed to supply only the basic load capacity with
peak load boilers in the district to supply additional peak load. There are enough examples of transport
grids that can transport the full heat demand thanks to the advantages of having all of the generation
facilities (main source as well as backup/peak load boilers) in one place. In the summer, however, when
the demand for heat is low, this does lead to low flow speeds in the heat transport grid because the latter is
dimensioned for the maximum load in the districts.
In grids that are supplied from several main heat sources, the utilization is optimized between the units
(dispatch). This is the case, for example, in cities such as Stockholm, Gteborg, and Vienna. In the current
projects in the Netherlands, this is hardly ever the case which is why it will not be discussed here.
5.4.3

Near Future Developments

The energy-saving potential of district heating, and the knowledge that our fossil fuel reserves are
limited, have increased the interest in district heating in the last years. In North Holland (Amsterdam
and environs), Randstad Zuid (Rotterdam and environs), the region of Utrecht and other areas such
as Almere and the KAN region (Arnhem-Nijmegen hub and environs), progress is being made with the
large-scale utilization of residual heat from large power stations, waste incinerators, and industry.
Besides the increasing interest technical developments in heat distribution systems are of influence on
the energy losses in the system and therefore the following developments are of importance:

110

Chapter 5 Transport and Distribution of Energy

Low temperature supply temperatures


Reducing the supply temperature has following economic advantages:
Reduced heat loss from the pipe network cause this loss is large and primarily dependent on
the temperature of the supply hot water (omakli et al, 2003, Poredo et al, 2001);
Efficiency at especially CHP plants is increased (better electricity efficiency because low-value
heat from the steam turbine(s) can be used);
More (industrial) lower temperature waste heat sources are suitable to deliver their heat to a
DH system.
In case the return temperature is unchanged, a lower supply temperature leads to a higher water mass
flow through the pipe at a given amount of thermal heat that has to be transported. The cost of pumping
operation therefore is increased. Pipe dimensions also have to increase, i.e higher cost, to avoid that the
physical speed limits in the grid are not exceeded. To ensure a minimum tap water temperature of 60C
and rule out any risk of Legionella pneumophila, supply temperatures in the transport grid can not be
decreased under the 65C. In case tap water is made with other heat sources, like heat pumps or with an
electrical boiler, the minimum supply temperature can be lowered based purely on the required
temperature levels for space heating only, which in the new build houses can go down to 40 - 50C.
Generating tap water with a heat pump or by electrical means of course other energy savings but in
combination with the advantages of a lower supply temperate in general can lead to more total energy
savings. This is not further investigated in this research.
Low temperature return temperature
Reducing of the return temperature in a DH system has also economic advantages (Ulbjerg et
al, 2005):
Reduced heat loss from the pipe network;
Cost for pumping operation are reduced;
Efficiency at especially CHP plants is increased (better electricity efficiency and better
condensation);
Efficiency of possible accumulation is increased in case supply temperature is not changed;
Pipe dimensions can be reduced in future renovations in case supply temperature is not
changed.
Development of pre-insulated pipes
There will be an increased demand for improved insulation properties in the future. Tendency towards
the use of a thicker layer of insulation and more efficient insulation materials than the polyurethane
foam (PUR) used today is the result (Srensen, 2001). Lower supplies temperatures to minimize the
heat loss can lead to a more widespread use of different plastic materials for DH pipes.
Plastic pipeline materials
For a long time, the industry has been looking for distribution systems that equal the quality of the
conventional Steel-PUR-PE system but do not have its disadvantages. The focus very quickly shifted
to plastic systems. In the past, pipes with a PEX-based medium pipe were laid on a large scale.
Because of the status of technology, this was not very successful neither technically nor cost-wise. A
new and promising development is a pipe system whose medium pipe is made of polybuthane, a PE
foam insulation, and an outer jacket of corrugated HDPE. Although the polybuthane medium pipe has
a higher expansion coefficient than steel, it has a much lower elastic modulus than steel meaning that
the expansion of the pipe due to temperature variations of the medium does not have to be taken into
account. Polybuthane can withstand bacteria better than PEX and other plastics. It also displays the
same behavior as steel. The big advantage of such a system is its enormous flexibility, which makes it
easy to lay pipes in the route's bends. The pipe is manufactured in long lengths (for example, up to
200 m for the smaller dimensions), which drastically reduces the number of connections (potential leak
locations) compared to the old system, which is made of 12-meter pipes and loose bends.
The area of application is limited to distribution grids with a low pressure (up to 6 bar) and a low
temperature (90C maximum). The pipes can be used to distribute central heating and hot tap water.
Moreover, these grids make up the lion's share of the distribution grid in terms of length and costs.
It is expected that using these plastic pipes will considerably reduce the construction and maintenance
costs of this type of distribution grid.

111

Part 2

Integral Pipe Tunnels


The Amsterdam Zuidas region is being developed on both sides of Amsterdam's A10 South ring road
between the Amstel and Schinkel rivers. This ambitious project includes building many new offices
and business facilities, and will be developed over a period of 25 years. Its favorable location in
relation to Schiphol Airport and the center of Amsterdam makes the Zuidas a highly attractive location
for businesses. To supply the region with energy, the Zuidas has an Integral Pipe Tunnel (IPT) in the
Mahlerlaan which contains all of the facilities including district heat transport pipes, district cooling
transport pipes, electricity cables, telecommunication cables, drinking water pipes, sewerage, and gas
pipes. The adjacent buildings are linked to these facilities from the tunnel. Although this may seem to
be a costly solution, its benefits lie in the accessibility and maintenance of the pipes in the tunnel. In
addition, the street does not have to be dug up to maintain existing pipes and cables or lay new ones,
which causes less inconvenience for traffic and residents.
5.4.4

District Heating and District Cooling

Most new office buildings, shopping centers, hotels etc. are equipped with air condition. Also in some
private homes, small local air-conditioning units are installed nowadays (Foged, 2003). Cooling is
therefore very relevant and will become more relevant in the near future to satisfy the growing request
for ideal working conditions in the shape of high thermal comfort. District Cooling (DC) is a cooling
system where absorbed thermal energy at consumers places, with use of a (cooling) water as
transport medium with supply temperatures of 6C and return temperatures of 16C, is transported
via a distribution network and cooled with chillers or a combination of free cooling and chillers to 6C
again. DC is an interesting complementary business for district heating companies because the
majority of the district heating grids are located nearby the consumers where the climate conditions
require heat in the winter and cooling in the summer to secure a high level of comfort that many
consumers request today. Advantages of DC are (Foged, 2003; Aastrup, 2003):
DC is interesting as an alternative to individual cooling equipment form the consumers sight;
Building owners can outsource secondary activities such as securing cooling and heat
delivery;
Large refrigeneration plants normally have higher efficiencies than individual cooling
equipment;
The coincidence factor between different consumers can often reduce the needed electrical
capacity of the production unit, and at the same time the reliability of the system is improved,
possibly in combination with accumulator tanks;
Different types of refrigeneration units can be used in the same system, making it easier to
optimize the operation of the entire DC system, for example district heat operating absorption
chillers for base load and compressor chillers for medium and peak load cooling;
Utilization of cheap sources (free cooling) like cold deep water form the sea (Stockholm,
Toronto) or deep lake water (Amsterdam) and surplus heat from industries/waste
incinerators/power plants to operate absoption chillers (Copenhagen, Gothenburg and
Helsinki);
Reduction in peak load during summer conditions in the electricity system in case conventional
air-condition are replaced by DC systems with different sources (free cooling, district heat and
electricity);
Cooling demand is present when the demand for heat is low, therefore increasing operating
times of consumption of district heat for cooling;
The combination of heat and cooling can make district energy feasible in areas where heat
alone is not feasible;
In case smaller refrigeneration units, based on absorption chillers which can operate on DH
with supply temperatures of 70C or lower, come available for private homes, office buildings,
shopping centers, hotels etc the existing DH network can be used during summer times to supply
heat for cooling.
A recently and first real commercial DC system, is the Nuon DC system in the Amsterdam Zuidas
which included a transport pipeline for lake water to the production facility or Cooling Plant, a
production facility with : lake water heat exchangers for free cooling, lake water heat exchangers for

112

Chapter 5 Transport and Distribution of Energy

chiller cooling, lake water pumps, DC pumps, other electrical equipment and a cooling distribution
network (see Figure 5.13).

Figure 5.13: Nuon District Cooling system Amsterdam Zuidas (Source : Nuon, 2006)
To generate the cooling Nuon is making use of free cooling from the Nieuwe Meer, a lake to the southwest of Amsterdam. Cold lake water, with a temperature of 5 - 6, is pumped form a depth of around
30 meters and is transported to the production facility. In the lake water heat exchangers for free
cooling, this water is cooling the return water, with a temperature of 16C, from the cooling
distribution network. At periods when the temperature in the lake is to high or the cooling demand is to
high for free cooling only, the chillers will secure a return temperature of 6C in the cooling
distribution network. For chiller cooling the lake water is also used but after the lake water has first
been used in the heat exchangers for free cooling as described above. After using the lake water, it
returns, with increased temperature, to the lake again. It is expected that the DC system, with a
designed capacity for peak cooling demand of 76 MWth, will have a Coefficiency Of Performance
(COP) of 10, which is compared with traditional chiller installations in building with COP of 2.5, four
times better (Euroheat & Power, 2006). A second DC system in the area of Bullewijk, located between
Schiphol Airport and the City of Amsterdam, will come in operation as of 2009. Here the free cooling is
taken form the Lake Ouderkerkerplas and the cooling capacity of the system will be almost equal as
the Zuidas system. The Zuidas region however is the first area in the Netherlands where both DH and
DC systems are available for the consumers.
5.5

Summary

The present day applied technical configuration for the transport and distribution grids for electricity,
gas, district heat and district cooling in the Netherlands are described. The main technological
developments for the future scenarios in 2012 and 2025 are also summarized in case these
developments are of importance for the Dutch electricity supply model.

113

References

References
Aastrup, B.(2003), Why is district cooling interesting business? Danish Board of District Heating, News
from DBDH 4/2003
Akzo Nobel, Edon, General Electric (1996), Delesto2, The Netherlands Most Efficient, Reliable and
Environmentally Compatible Combined-Cycle Cogeneration Plant, Brochure
Bahrman, M., et al. (2007), The ABCs of HVDC Transmission Technologies: an overview of high
voltage direct current systems and application, IEEE Power & Energy Magazine March/April 2007
Vol.5 No.2
Business Highlights (1999), ELSTA Power Plant, voor milieuvriendelijke energie [Elsta Power Plant,
for Environmentally-Friendly Energy]
Castanier, A., Beau, J.-C. (1988), Power Plant Repowering by Steam-Gas Combined Cycle
Conversion, Alsthom Review No.10 - 1988
CBS / Cogen Project (2004), Update gasmotorenbestand CBS [Update Gas Engine Files CBS]
CBS (2006), Duurzame energie in Nederland 2005, ISBN 978-90-357-1896-8 [Sustainable Energy in
the Netherlands 2005]
CBS (2007), Duurzame energie in Nederland 2006, ISBN 978-90-357-1518-9 [Sustainable Energy in
the Netherlands 2006]
Cogen Projects (2004), Belichten met elektriciteit uit eigen WKK of inkopen op de vrije markt,
Document CP 03.136 EK [Lighting with electricity from our own CHP or bought in on the free market?]
Cogen Projects (2005), Voortgangsrapportage 2005 Warmte/Kracht in de Glastuinbouw, uitgave 1
[Progress Report 2005 on CHP in the Market Garden Sector]
omakli, K., Yksel, B., omakli, O. (2003), Evaluation of Energy and Exergy Losses in District
Heating Networks, Applied Thermal Engineering 24 (2004) 1009 - 1017
Correlj, A., Van der Linde, C., Westerwoudt, T. (2003), Natural Gas in the Netherlands, From
Cooperation to Competition?, Oranje-Nassau Groep, ISBN 90 5031 0842
De Jong, H. (2003), Decentraal vermogen een kansrijke optie, 2003, MSc Thesis, Delft University of
Technology [Decentralized Capacity: an opportunity-rich option]
Dondi, P., et al. (2002), Network Integration of Distributed Power Generation, Journal of Power
Sources 106 (2002) 1 - 9
ECN (2003), Kosten Duurzame Elektriciteit, Kleinschalige waterkracht, ECN-C--03-074/G [Costs of
Sustainable Electricity, Small-Scale Hydro-Power]
Electrabel (2000), Centrale Gelderland, Brochure [Gelderland Power Plant]
Electriciteits-Productiemaatschappij Oost- en Noord-Nederland EPON (1980s), Excursieroute
Centrale Gelderland [Excursion Route Gelderland Power Plant]
Electriciteits-Productiemaatschappij Oost- en Noord-Nederland EPON (1980s), Eemscentrale,
Brochure [Eems Power Plant]
Electriciteits-Productiemaatschappij Oost- en Noord-Nederland EPON (1980s), Warmtekrachtcentrale
Almere, Brochure [Almere Cogeneration Plant]

114

References

Electriciteits-Productiemaatschappij Oost- en Noord-Nederland EPON (1980s), Almere Cogeneration


Plant, Brochure
Energieproductiebedrijf UNA (1996), Purmerend Combined Heat & Power Plant, Brochure
Energy Delta Institute (2005), De Gaswaardeketen, Course [The Gas Value Chain]
Energietechniek (1999), Flexibele stoomproducent, No. 11, Volume 77 [Flexible Steam Producer]
Energietechniek (1996), Demkolec op kruispunt, No. 11 jaargang 74, 692 695 [Demkolec at a
Crossroads]
Energietechniek (1996), Spanningsveld rond schone kolen, No. 11 jaargang 74, 690 691 [Areas of
Tension around Clean Coal]
Energietechniek (1997), Spannende tijden voor PER+, No. 2 jaargang 75, 100 101 [Anxious Times
for PER+]
Energietechniek (1997), Flexibele energiecentrale voor PER+, No. 2 jaargang 75, 102 105 [Flexible
Power Plant for PER+]
Energietechniek (1998), Stap voor stap naar wkk-installatie, No. 9 jaargang 76, 474 478 [Step by
Step toward CHP Installation]
Elsta (2004), Publieks-milieujaarverslag 2003 [Public environmental report 2003]
E.on Netz (2003), Wind Power in Germany: present situation and outlook
EPRI (2006), DC Power Production, Delivery and Utilization, an EPRI White Paper
EPZ (1996), Warmte / krachtcentrale Moerdijk: energie op maat, Brochure [Moerdijk CHP Plant: tailormade energy]
EPZ (2000), Radioactief afval: isoleren, beheersen en controleren, Brochures [Radioactive Waste:
isolation, management and control]
EPZ (2004), Milieu overheids jaarverslag Koleneenheid Borssele 12 [Annual government
environmental report on Borssele 12 Coal-Fired Power Plant]
Essent Energie (2001), Grootschalige opwekking van elektriciteit en warmte, Brochure [Large-Scale
Generation of Electricity and Heat]
Essent (2005), Overheidsmilieujaarverslag 2004, Beschrijvende deel WKC Swentibold, [Annual
government environmental report 2004, Descriptive part CHP Swentibold]
Euoheat & Power (2006), District Cooling: cooling more with less
European Commission Energy (1999), Combination of a Waste Incineration Plant and a Combined
Cycle Power Plant
Foged, M. (2003), District Cooling a business opportunity, Danish Board of District Heating, News
from DBDH 4/2003
Fris, S., Eurlings, J. (2003), CHP for Industry, Articles series Cogeneration and On-Site Power
Production May 01, 2003
German WindEnergy Association (BWE) 2008, Status and Perspectives of the German Wind Energy
Market, WIREC 2008 Washington
Gas Transport Service B.V. (2007), Annual Report 2006

115

References

Gas Transport Service B.V. (2008), Beschrijving Gaskwaliteitssysteem, Zomer 2008, LNM 08.0231
[Description Natural Gas Quality System, Summer 2008]
GE Jenbacher (2005), Jenbacher Type 4, Gas Engine Product Brochure
Hotchkiss, R. (2002), Waste/Biomass Co-gasification with Coal, Report No. COAL R216 DTI/Pub URN
02/867
Huisvuilcentrale HVC (2006), Volumes in Beeld, www.huisvuilcentrale.nl/kerngegevens [Volumes
Mapped Out]
International Energy Agency IEA (2002), Distributed Generation in Liberalized Electricity Markets,
ISBN 9264198024
Interview with C. Smies (2006), Nuon Warmte
Interview with Professor G. van Kuijk (2005), TU Delft Lucht en Ruimtevaart, afdeling Windenergie
Interview with B. Dijkman (2005), Nuon Power Generation
Kehlhofer, R., Bachmann, R., Nielsen, H., Warner, J. (1999), Combined-Cycle Gas & Steam Turbine
Power Plants, Second Edition, PennWell, Tulsa, Oklahoma
Korobitsyn, M.A. (1998), New and Advanced Energy Conversion Technologies: analysis of
cogeneration, combined and integrated cycles, PhD thesis, Twente University of Technology
Kostic, D., Reinders, E. (1998), Stap voor stap naar wkk-installatie, Energietechniek 9, Volume 76,
September 1998 [Step by Step toward CHP Installation]
Ministerie van VROM (2004), Voortgangsrapportage Landelijk afvalbeheerplan
VROM/DGM/SAS BAOO May 2004 [Progress Report on National Waste Management Plan]

(LAP),

NAM & Gasunie (2001), Position Paper Ondergrondse Bergingen [Position Paper on Underground
Storage]
Novem (2001), Meestoken van biomassa in kolen gestookte E-centrales, Report no. 2EWAB01.30
[Biomass Co-Firing in Coal-Fired Power Plants]
Nuon (2006), District Cooling in Amsterdams Zuidas, Brochure
Nuon (2006), Gasunie en Nuon vergroten leveringszekerheid Nederlandse gasmarkt, Pres Release 13
[Gasunie and Nuon increase security of supply on the Dutch Gas Market]
Nuon (2006), Nuon Magnum, Multi-fuel Power Plant, Brochure
Nuon / CCE (2006), Showcase of District Cooling Systems in Europe Amsterdam, IEA DHC/CHP
case studies
N.V. Energieproduktiebedrijf UNA (1996), Warmtekrachtcentrale Purmerend, Brochure [Purmerend
Cogeneration Plant]
N.V. Nederlandse Gasunie (2005), Facts 2005, Appendix by N.V, Nederlandse Gasunie Annual
Report 2004
stergaard, P. (2004), Modelling Grid Losses and the Geographic Distribution of Electricity
Generation, Renewable Energy 30 (2005), 977- 987
Pepermans, G., et al. (2005), Distributed Generation: definition, benefits and issues, Energy Policy 33
(2005) 787 - 798

116

References

Pauwels, JP., et al (2000), Rapport van de Commissie voor de analyse van de Productiemiddelen van
Elektriciteit en de Reorientatie van de Energievectoren (Ampere) aan de Staatssecretaris voor Energie
en duurzame Ontwikkeling, Belgium [Report by the Commission for the analysis of the means of
electricity production and the reorientation of the energy vectors (Ampere) to the State Secretary for
Energy and Sustainable Development]
Poredo, A., Kitanovski, A. (2001), Exergy Loss as a Basis for the Price of Thermal Energy, Energy
Conversion and Management 43 (2002) 2163 - 2173
Provinciale Gelderse Energie Maatschappij PGEM (1970s), Beschrijving excursieroute Centrale
Gelderland [Excursion Route Gelderland Power Plant]
Ratafia-Brown, J, et al. (2002), An Environmental Assessment of IGCC Power Systems, Presented at
the 19th Annual Pittsburgh Coal Conference, September 23 - 27, 2002
SBB Schelde Breda Boilers B.V. (1990), Cogeneration Plant Delesto Location : Akzo - Delesto,
Delfzijl Holland, technical information document
Skog, J-E, et al. (2006), The NorNed HVDC Cable link, A Power Transmission Highway between
Norway and the Netherlands
Skog, J-E (2006), Interview with J-E Skog, Power Engineering International
Srensen, C.F. (2001), The Development of Pre-insulated pipes, Danish Board of District Heating,
News from DBDH 2/2001
Statnett (2004), Preparation for 1200 MW transmission capacity across Waddensea?, Memo 38053
Rev1
Statnett (2004), Comments regarding the NorNed Project, Letter to Office of Energy Regulation (DTE,
27.10-2004, reference 04/46-11
Stork Ketels (1995), CHP Eerbeek The Netherlands Combined Heat and Power Plant, Brochure
TenneT (2002), Verwacht veilig beschikbare landsgrensoverschrijdende transportcapaciteit voor
elektriciteit in het jaar 2003 in Nederland [Expected safely available Interconnector Capacity for the
year 2003 in the netherlands]
TenneT (2002), Capaciteitsplan 2003 2009 [Capacity Plan 2003-2009]
TenneT (2004), Toelichting import en export van elektriciteit, Notitie CA 04-0056 [Notes on the Import
and Export of Electricity]
TenneT (2004), NorNed HVDC Project Technische beschrijving [Technical Description of the NorNed
HVDC Project]
TenneT (2007), NorNed Europease schakel voor de toekomst, Brochure [NorNed European Link for
the Future]
TenneT / National Grid (2007), BritNed Europes link for the future, First electricity link between UK
and the Netherlands, Brochure
Thomassen International (1996), Trainingshandleiding cursus : Thomassen General Electric Heavy
Duty Gas Turbines
UCTE (2008), UCTE Transmission Development Plan, Edition 2008
Ulbjerg, F., Steffensen, H. (2005), Low return temperature can lead to big savings, Danish Board of
District Heating, News from DBDH 1/2005

117

References

Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2004), Trading between countries improving the balance
between Environment, Economy and Security of Supply or power play? The EU situation, Paper
presented at 19th World Energy Congress, Sydney, Australia, September 5-9, 2004
Van Eck, T. (2007), A New Balance for the Energy Sector: no longer a puppet in the hands of
technology, public interests and markets, ISBN 9789078889014
Vereniging Afvalbedrijven (2004), Energie uit afval, Statusrapportage 2004 [Energy from Waste,
Status Report 2004]
Vereniging Afvalbedrijven (2004), Afvalverwerking in Nederland, Gegevens 2003, Rapportnummer
AOO 2004-11 [Waste Processing in the Netherlands. Data 2003]
World Coal Institute (2003), Clean Coal - Building a Future through Technology, Report by the World
Coal Institute on the contribution of coal to global sustainable development.

Internet Addresses
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

118

http://www.districtenergy.com
http://www.electrabel.nl
http://www.elstacogen.nl
http://www.energyvalley.nl/
http://www.home01.wxs.nl/~windsh/
http://www.huisvuilcentrale.nl/kerngegevens
http://www.iea-dhc.org/index.html
http://www.iea.org/Textbase/stats/
http://www.ei.tbm.tudelft.nl
http://www.nationalgrid.com/uk/Interconnectors/Netherlands
http://www.senternovem.nl/KEI/21_netwerken/51_iea/11_end-use/el01_iea_dsm_start.asp
http://www.tennet.org
http://www.tennet.org/projecten
http://dsm.iea.org
http://ec.europa.eu/eurostat
http://ec.europa.eu/energy/res/sectors/bioenergy_incineration_en.htm

Part 3

Supply Trends

120

Chapter 6 Technological Trends in Supply Options

Chapter 6
6

Technological Trends in Supply Options


6.1

Introduction

This section traces the main technological developments in the electricity production options
discussed in Sections 4.4, 4.5 and 4.6. These developments serve as direct input for the model of the
Dutch electricity supply which will be applied for 2012 and 2025 and for the choice of product options
in the scenarios described in Part 4. First analyzed is conventional and combined cycle technology
(clean coal technology, nuclear technology, gas turbine and combined cycle technology) and the
expected developments in process characteristics, emissions, operational and maintenance costs, unit
size, efficiency, and levels of investment. Then follows a briefly focus on anticipated developments in
CHP technology based on gas turbines (already touched on in the discussion of conventional and
combined cycle technology), gas engines and micro CHP, paying specific attention to fuel, emissions,
operational and maintenance costs, and efficiency. The developments in sustainable production
options will then be outlined insofar as they are relevant to the model (see the bibliography for further
information). This also applies to information on developments in other technologies such as geothermal storage, heat pumps, fuel cells and cooling, and the technological developments in energyconsuming equipment, processes and buildings which are not directly relevant to the model. The
future key points relating to conventional and combined cycle, CHP and the sustainable options will
then be summarized in tables as input for the model.
6.2

Conventional and Combined Cycle Technology

When we described the technology of the conventional and combined cycle units in Section 4.4 of
Chapter 4 a division was made into conventional technology, nuclear technology and gas turbine and
combined cycle technology. In the next sections the main technological developments for these
electricity production options is given with a focus on developments in process characteristics, fuel
and emissions, operation and maintenance, unit size, net electrical efficiency1, and investment levels.
6.2.1

Clean Coal Technologies (CCTs)

Clean Coal Technologies (CCTs) facilitate the use of coal in an environmentally satisfactory and
economically viable way. Among other aspects, they meet various regulations covering emissions,
effluents, and residues. In some situations, CCTs offer the possibility of satisfying even more stringent
standards, at an acceptable cost (IEA Clean Coal Centre, World Coal Institute)
A basic approach to the cleaner use of coal is to reduce emissions by reducing the formation of
pollutants such as NOx and/or cleaning the flue gases after combustion. A parallel approach is to
develop more efficient systems so that less coal is used to generate the same amount of power,
together with improved techniques for flue gas cleaning, for effluent treatment, and for residues use or
disposal. Electrical efficiency may also be increased by using a higher grade coal.
Various technologies are undergoing development in order to provide an environmentally satisfactory
method of using coal as a basic fuel for power production in new plants. All methods which meet
environmental standards in different parts of the world are included. Some are now commercially
available, backed by large-scale operating experience in a number of countries. Others are still at the
demonstration stage. These technologies include:
1

In this section the mentioned net electrical efficiencies (LHV) for new technology power plants are based on
solely electricity production. However, these power plants can also be designed to deliver process and/or district
heat. In that case distinguish can be made into electrical and thermal efficiency (see Section 8.4.2 of Chapter 8).

121

Part 3

Pulverized Coal Combustion (PCC) with supercritical steam driving a steam turbine, together with
flue gas cleaning units.
Atmospheric pressure fluidized bed combustion (FBC) in both bubbling (BFBC) and circulating
(CFBC) beds, mainly with sub critical steam turbines, together with sorbent injection for SO2
reduction and particulates removal from flue gases.
Integrated gasification combined cycle (IGCC), using different types of gasifier, and in combined
cycle with both a gas and steam turbine. The syngas stream is cleaned of H2S and particulates,
before combustion and expansion of the combustion products through the turbine. Various levels
of integration are used (see Section 6.2.3).
Carbon capture & storage (see Section 6.2.5).

Next-generation clean coal technologies include Hydrogen from coal, integrated gasification fuel cells
and underground coal gasification. Research in this technologies continues but this is beyond the
scope of this reserach as it is assumed that large scale implementation is not likely within the next 30
years. Here we concentrate on the development of the five above mentioned technologies which have
a potential for large scale applications within the next 30 years as they are crucial for the electricity
supply model (see Part 4). For a detailed explanation of the next-generation clean coal technologies
see World Coal Institute, IEA Clean Coal Centre and other relevant literature.
Pulverized Coal Combustion
Pulverized Coal Combustion (PCC) is the most commonly used method in coal-fired power plants, and
is based on many decades of experience. Units operate at close to atmospheric pressure within the
boiler, simplifying the passage of materials through the plant. PCC can be used to fire a wide variety of
coal, lignite and biomass, although it is not always appropriate for those with a high ash content.
Process Characteristics
Pulverized Coal Combustion units use a Rankine Cycle to convert the thermal energy of the steam
into mechanical energy. Process improvements of the past decade involves the use of superheat and
singular reheat sections, the use of sub critical or supercritical pressures, see also Section 4.4.1.
Two broadly different boiler designs are used. One is the traditional two-pass layout where there is a
furnace chamber, topped by some heat transfer tubing to reduce the flue exit gas temperature
(FEGT). The flue gases then turn through 180, and pass downwards through the main heat transfer
and economizer sections. The other design is to use a tower boiler, where virtually all the heat transfer
sections are mounted vertically above each other, over the combustion chamber.
Most PCC boilers operate with what is called a dry bottom. Combustion temperatures (with bituminous
coal) are held at 1500 - 1700C. With lower rank coals the range is 1300 - 1600C. Boilers which use
anthracite as fuel commonly use the downshot burner arrangement to achieve longer residence times
and ensure carbon burn-out. Downshot burners send the coal-air mixture down into the cone at the
base of the boiler.
The principal future developments for PCC units involve (Bugge et al, 2003; Hellfritsch et al, 2007;
Kitto,1996):
Increasing net plant efficiencies by raising the (supercritical) steam pressure (see Figure 6.1)
and steam temperature at the boiler outlet/steam turbine inlet with the now under development
new high temperature nickel-based superalloys (Bugge et al, 2003).
Double reheat design instead of single reheat.
Oxyfuel technology which means burning the coal with pure oxygen instead of air.
Improvements of steam turbine efficiency through steam flow path design improvements made
possible by the latest 3 dimensional numerical modeling.
Reduction in auxiliary power loads at full loads and part loads, particularly in the environmental
equipment area.
Ensuring that units can load following satisfactorily.
Ensuring that flue gas cleaning units can meet future emissions limits and environmental
requirements.

122

Chapter 6 Technological Trends in Supply Options

60

Net Electrical Efficiency [%

50

40

30

20

10

0
1

Pressure [bar]

180

240

240

310

310

400

350 - 400

SH Temp [C]

540

540

540

540

595

650

700 - 720

RH Temp [C]

540

540

550

550

565

595

~ 700

2RH Temp [C]

--

--

565

565

565

595

~ 700

Figure 6.1: Effect of increasing steam pressure and temperature on net electrical plant efficiencies for
PCC units (Source: Kitto, 1996; Bugge et al, 2003)
Flue gas cleaning and solid removal
With appropriate flue gas cleaning units emissions (NOx, SO2) and solid removal from new PCC units
can meet all current requirements reliably and economically, using well-proven technology. The
necessary emission control measures can be taken with a relatively small effect on the net plant
efficiency, although the capital cost of these measures can represent up to one third of the cost of the
unit when meeting the most stringent current standards. For a detailed explanation of these
technologies see IEA Clean Coal Centre, World Coal Institute (2003), Botha (2004), DTI (2000) and
other relevant literature.
Operation and Maintenance
The operation and maintenance (O&M) cost covers on- and off site personnel, maintenance materials,
supplies and expenses, regulatory fees, and insurances (Boer, 2004). Part of these cost (on-site
personnel, regulatory fees, insurances) can be considered fixed cost, regardless of the units operation
time at least if the unit is not mothballed; another part (off-site personnel, maintenance materials) is
variable cost depending on the units operation time. In most plants, planned maintenance procedures
are adopted relative to the number of operating hours and a few use breakdown maintenance.
Condition-based preventive maintenance, which relies on the determination of the condition of the
component for the purpose of its repair and replacement is the newest strategy. This can sometimes
extend the life of a component well beyond fleet average (IEA Clean Coal Centre).
As stated above the operation and maintenance (O&M) cost consists of a fixed and a variable part.
The variable part is of importance to the model as it impacts the calculated dispatch cost. Coal-fired
units have a relatively large share of fixed cost the operation of a large unit nowadays requires 6
staff working in shifts and this is a factor 3 more compared to a gas-fired plant with 2 staff working2.
The variable maintenance costs are lower and usually the unit is taken out of operation once every
two years for a longer inspection. In VGB report CO2 Capture and Storage (2004), O&M cost for a
2

Today trends of cost savings are leading to concentration of staff working on mainly (large) locations on which
also the control of relatively smaller gas fired units takes place.

123

Part 3

500 MW PCC unit is 7.0/MWh. In a study of the Uranium Information Centre (2006) in which the
economic of the nuclear generation is compared with coal fired generation and other generation
capacity, O&M costs for coal fired generation is set on 7.4/MWh. An older ECN study in
collaboration with Kema (ECN, 1995) also provides estimates of the fixed and variable costs for future
production units. Though the absolute amounts no longer apply, it can be concluded from this study
that the fixed O&M costs represent about 70% of total cost and the variable O&M costs approximately
30%. Excluding inflation, amongst others, no changes are expected in both absolute amounts and the
ratio between fixed and variable in the period between now and 2030 (ECN, 1995). For the model
calculations a fixed cost of 5.0/MWh and variable cost of 2.0/MWh is used for 2012 and 2025.
Unit Size
PCC boilers are built to match steam turbines which have outputs between 50 and 1,300 MWe. In
order to take advantage of the economies of scale, most new units are rated at over 300 MWe, but
there are relatively few really large ones with outputs from a single boiler/turbine combination of over
700 MWe. This is because of the substantial effects such units have on the distribution system if they
should trip out for any reason. For the model calculations 1,200 MWe units (2 blocks) are used.
Net Electrical Efficiency (based on LHV)
One of the driving forces which is currently encouraging the use of more efficient power plants is the
environmental concern in many countries, and the declared goal of most OECD governments to
reduce CO2 emissions to 1990 levels according to the Kyoto protocol. This is a goal which leaves
power generators with many unsolved problems, but increasing the electrical efficiency of converting
coal to power is one of the less expensive ways of reducing CO2 emissions. It does, however, involve
the construction of improved boilers and turbines designs, as the costs for retrofitting an existing sub
critical boiler to a supercritical steam system would be prohibitive.
Increasing the net electrical efficiency has the potential to reduce other emissions per MWh
generated, such as those of SO2 and NOx. Where the coal cost is high, as where traded coals are
used, increasing net electrical efficiency can result in reduced overall costs in new plants for power
generation, as less fuel is needed. The average net electrical efficiency of some older, smaller units
burning, possibly, poor quality coals can be as low as 30%. A commonly used assumption for the
average net electrical efficiency of larger existing plants with sub critical steam burning somewhat
higher quality coals is that it is in the region of 36 - 38%. Newer plants (commissioned between 1990 2000), however, with supercritical steam can now achieve net efficiencies in the 42 - 47% range.
New high temperature alloys are under development with the aim of facilitating steam temperatures as
high as 700C (Bugge, 2003). This could make net electrical efficiencies of 52 - 55% achievable with
PCC (see Figure 6.1). According to Siemens (Voges, 2004) application of new advanced materials in
combination with ultra critical steam conditions to PCC power plants should enable electrical
efficiencies of 53% whilst the Australian Coal Association mentioned 55% to be achieved in the future.
Gasteiger (2006) is expecting a target electrical efficiency of > 52% in 2020 for lignite-fired steam
power plants. With a 2 - 3% higher electrical efficiency of coal-fired steam plants compared with lignite
(situation as of 2005), this would mean a target efficiency of > 54% in 2020 for coal-fired power plants.
For the model calculations a (base load) net electrical efficiency of 45% is used for 2012 and 53% for
2025.
Investment3
In VGB Report (2004) a specific investment for a 500 MW PCC unit is given of 1,029/kWel. In a CE
study (2005) for a 1000 MWe PCC unit an specific investment of 1,200/kWel is applied. E.on
Benelux will invest approximately 1.2 billion to build a 1100 MWe hard coal fired power plant next to
their existing location on Maasvlakte, due to start in 2009/2010 (E.on, 2006). The total specific price is
1,090/kWe including other cost than EPC but with the remark that probably facilities of the existing
units can be used. According to Blankenship (2006) the partners GE and Bechtel assume a 600 MW
3

Since early 2007, investment price levels have gone up substantially caused by the high tension on the electric
power plant construction market. Because this is applicable for all production options only the absolute investment
levels have increased compared with the price levels used in this thesis. The relative differences in price levels
however, remains the same.

124

Chapter 6 Technological Trends in Supply Options

supercritical PCC plant in Ohio Valley would cost $ 1,200/kW at a brownfield site and $ 1,460/kWe at
a greenfield site. It is likely that these costs are exclusive of other than EPC cost.
For the model calculations the investment cost for a PCC are set at 1,200/kWe in 2012 decreasing
to 1,100/kWe in 2025.
Fluidized Bed Combustion (FBC)
Fluidized bed combustion technologies include:
Atmospheric pressure fluidized bed combustion (FBC) in both bubbling (BFBC) and circulating
(CFBC) beds, mainly with sub critical steam turbines, together with sorbent injection for SO2
reduction and particulates removal from flue gases.
Pressurized fluidized bed combustion (PFBC) mainly using bubbling beds, and in combined
cycle with both a gas and steam turbine. Sorbent injection is used for SO2 reduction and
particulates removal from flue gases. Pressurized circulating fluidized bed combustion
(PCFBC) is currently being demonstrated.
The coal-fired power generation capacity using FBC showed rapid growth between 1985 and 1995,
but it still represents less than 2% of the world total (IEA Clean Coal Centre). Technology of the
Fluidized Bed Combustion is therefore, and the assumption that large scale implementation is not
likely within the next 30 years, not used for the model calculations. However atmospheric pressure
fluidized bed combustion is short described in the next part to point out the major differences with
PCC. For a more detailed description of Fluidized Bed Combustion (atmospheric and/or pressurized)
please refer to the literature.
Bubbling Fluidized Bed Combustion (BFBC) at Atmospheric Pressure
FBC in boilers at atmospheric pressure can be particularly useful for high ash coals and/or those with
variable characteristics. The combustion chamber consists of a bed of relatively dense and coarse
inert particles of around 3 mm in size to which the coal is continuously fed. The combustion air is fed
from the bottom of the bed and it keeps the solid inventory in a fluidized state. The thermal inertia of
the large mass of hot particles allow coarse coal or lignite particles with high ash contents to be fully
converted (large residence time).Two formats are used, bubbling beds (BFBC) and circulating beds
(CFBC).
Process Characteristics
Combustion takes place at temperatures between 800 and 900C. Bubbling beds use a low fluidizing
velocity, so that the particles are held mainly in a bed which will have a depth of about 1 m, and has a
definable surface. Sand is often used to improve bed stability, together with limestone for SO2
absorption. As the coal particles are burned away and become smaller, they are elutriated with the
gases, and subsequently removed as fly ash. In-bed tubes are used to control the bed temperature
and generate steam. The flue gases are normally cleaned using a cyclone, and then pass through
further heat exchangers, raising steam.
Emissions and Residues
See section on CFBC.
Operation and Maintenance
See section on PCC.
Unit Size
Atmospheric BFBC is mainly used for boilers up to about 25 MWe, although there are a few larger
plants where it has been used to retrofit an existing unit. There are hundreds of small BFBC units in
China.
Net Electrical Efficiency (based on LHV)
The net electrical efficiency is around 30%.

125

Part 3

Circulating Fluidized Bed Combustion (CFBC) at Atmospheric Pressure


Process Characteristics
Circulating beds use a higher fluidizing velocity, so the particles are constantly held in the flue gases,
and pass through the main combustion chamber and into a cyclone, from which the larger particles
are extracted and returned to the combustion chamber. Individual particles may recycle anything from
10 to 50 times, depending on their size, and how quickly the char burns away. Combustion conditions
are relatively uniform through the combustor, although the bed is somewhat denser near the bottom of
the combustion chamber. There is a great deal of mixing, and residence time during one pass is very
short. The direct injection of limestone into the bed offers the possibility of economic SO2 removal
without the need for flue gas desulphurisation.
The design must take into account ash quantities, and ash properties. While combustion temperatures
are low enough to allow much of the mineral matter to retain its original properties, particle surface
temperatures can be as much as 200C above the nominal bed temperature. If any softening takes
place on the surface of either the mineral matter or the sorbent, then there is a risk of agglomeration or
of fouling.
Emissions and Residues
Combustion takes place at temperatures between 800 and 900C resulting in reduced NOx formation
compared with PCC. N2O formation (296 GWP, IPCC (2001)) is, however, increased. SO2 emissions
can be reduced by the injection of sorbent into the bed, and the subsequent removal of ash together
with reacted sorbent. Limestone or dolomite are commonly used for this purpose. Mostly only
particulates removal is required in order to meet emissions limits, but in some places NOx reduction is
also required. Carbon-in-ash levels are higher in FBC residues than in those from PCC.
Operation and Maintenance
See section on PCC.
Unit Size
Atmospheric CFBC is used in a number of units of around 250 - 300 MWe, and there are a number of
commercially operating plants. There are designs for units up to 600 MWe. A 460 MWe supercritical
unit is under construction at Lagisza, Poland (situation as of 2007).
CFBC boilers represent the market for relatively small units, in terms of utility requirements. They are
used more extensively by industrial and commercial operators in smaller sizes, both for the production
of process heat, and for on-site power supply. A few are used by independent power producers,
mainly in sizes in the 50 to 100 MWe range.
Net Electrical Efficiency (based on LHV)
In the 100 - 200 MWe range, the net electrical efficiency of FBC units is commonly a little lower than
that for equivalent size PCC units by 3 to 4 percentage points. In CFBC, the heat losses from the
cyclone/s are considerable. This results in reduced efficiency, and even with ash heat recovery
systems, there tend to be high heat losses associated with the removal of both ash and spent sorbent
from the system. The use of a low grade coal with variable characteristics tends to result in lower
efficiency, and the addition of sorbent and subsequent removal with the ash results in heat losses. The
supercritical unit at Lagisza is expected to have a net electrical efficiency of over 40%.
6.2.2

Nuclear Technologies

Nuclear power is an important source of electric in many countries of the world. Despite a declining
share of global electricity production, nuclear power is projected to remain an important source of
electric power through 2030. Electricity generation by Nuclear power plants around the world in 2003
was 2,523 TWh which, on a total world net electricity consumption in 2003 of 14,781 TWh means a
share of 17% (EIA, 2006). Nuclear technology is constantly being improved, reactor designs are more
and more inherent safe, and total plant efficiencies are increasing. Figure 6.2 shows reactor
development over the past decades.

126

Chapter 6 Technological Trends in Supply Options

Whereas nowadays third-generation reactors are installed in the near future (2010 - 2040) reactors of
Generations III+ and IV will be introduced. For an extensive description of all reactor designs for the
near future please refer to U.S. DOE NERAC (2002).

Figure 6.2: Reactor development over the past decades and future prospects (Source: U.S. DOE
NERAC, 2002)
In recent years, the Nuclear Regulatory Commission (NRC) has set up a process by which reactor
designs might be certified prior to any actual construction plans, see Table 6.1. Any new reactor built
in the United States over the next decade or so would most likely use designs either recently certified
by the NRC or that will be certified by the NRC in the near future (design approval can alternatively
coincide with construction and operation licensing, skipping the certification process). The re-creation
of older designs is popular overseas and cannot be ruled out in the United States.
Reactor
Design

Vendor

Approximate
Capacity (MWe)

Reactor Type

Certification Status

Target
Certification

AP600

Westinghouse

650

PWR

Certified

Certified

AP1000

Westinghouse

1,117

PWR

Certified

Certified

ABWR

GE et al

1,371

BWR

Certified

Certified

Westinghouse

1,300

PWR

Certified

Certified

GE

1,550

BWR

Undergoing certification

2007

AREVA NP

1,600

PWR

Pre-certification

2009

Eskom

180

HTGR

Pre-certification

Not Available

Westinghouse et al

360

PWR

Pre-certification

2010

Mitsubishi

1,600

PWR

Undergoing certification

2011

Pre-certification

Not Available

System 80+
ESBWR
EPR

Westinghouse,
PBMR
IRIS
US APWR

Modified
ACR Series

AECL

700-1,200

PHWR

Research prototype
GT-MHR
4S

General Atomics

325

HTGR

planned

Not Available

Toshiba

10 - 50

Sodium-cooled

Potential construction

Not Available

Table 6.1: Certification process for new reactors in the US (Source: EIA, Release date : November
2006)

127

Part 3

Presently there are four certified new reactor designs in the United States: the System 80+, the
Advanced Boiling Water Reactor (ABWR), the AP600 and the AP1000. These designs are sometimes
called Advanced Light Water Reactors (ALWR) because they incorporate more advanced safety
concepts than the reactors previously offered by vendors. They are also sometimes called Generation
III reactors to distinguish them from earlier designs now operating in the US and globally and from
later designs now seeking certification which are sometimes called Generation III plus. Design
certifications can expire if not supported by a vendor.
For this research, focus is on the ABWR, EPR and PBMR (to be explained below) as being certified
(ABWR) and as pre-certification design (EPR and PBMR) and possible options for future reactors in
the Netherlands.
Advanced Boiling Water Reactor (ABWR)
The Advanced Boiling Water Reactor (ABWR) is an improved design of the boiling water reactor
(BWR) and the only design, among the third generation designs, with construction and operating
experience. The ABWR design was developed by a consortium led by Toshiba, GE and others
(Toshiba Corporation et al, 2005). The ABWR design is already licensed in three countries: the United
States, Japan and Taiwan. Meanwhile the next generation ABWRs (Generation III+ ) are now being
developed with a passive safety system among other design features to improve economic
competitiveness (IAEA, 2006; Murase, 2005).
Main Components and Processes
In Figure 6.3 the main components of an ABWR are shown followed by a brief description of two of the
main components (Toshiba Corporation, 2007).

Figure 6.3: Main components of ABWR (Source: Toshiba Corporation, 2007))

128

Chapter 6 Technological Trends in Supply Options

Reactor Pressure Vessel and Its Internals


In an ABWR, coolant water is re-circulated inside the reactor pressure vessel (RPV) by reactor internal
pumps installed within the vessel itself, see Figure 6.3. The RPV contains the fuel assemblies, control
rods and the reactor's internals. It is made of low-alloy steel and its inner surface is clad with
corrosion-resistant materials. The internal elements it houses include the steam/water separator and
dryer and the core support structures.
Primary Containment Vessel
The primary containment vessel encloses the reactor pressure vessel, other primary components and
piping. In the highly unlikely event of an accident, this shielding prevents the release of radioactive
substances. The ABWR uses a Reinforced Concrete Containment Vessel (RCCV). Its reinforced
concrete outer shell is designed to resist pressure, while the internal steel liner ensures the RCCV is
leak-proof. The compact cylindrical RCCV integrated into the reactor building enjoys the advantages
of earthquake-resistant design and economic construction cost.
Safety System
The ABWR has three completely independent and redundant safety systems. The systems are
mechanically separated and have no cross connections as in earlier BWRs. They are electronically
separated so that each safety system has access to redundant sources of AC power and, for added
safety, its own dedicated emergency diesel generator. Each safety system is physically separated and
is located in a different quadrant of the reactor building, separated by fire walls. A fire, flood or loss of
power which disables one safety system has no effect on the capability of the other safety systems.
Finally, each safety system contains both a high and low pressure system and each system has its
own dedicated heat exchanger to control core cooling and remove decay heat. One of the high
pressure systems, the Reactor Core Isolation Cooling (RCIC) system, is powered by reactor steam
and provides the diverse protection needed should there be a station blackout.
The safety systems are capable of keeping the core covered with water at all times. Because of this
capability and the generous thermal margins built into the fuel designs, the frequency of transients
which will lead to a scram and therefore to plant shutdown has been greatly reduced (to less than one
per year). In the event of a loss of coolant accident, plant response has been fully automated and
operator action is not required for 72 hours, the same capability as for the passive plants.
Operation and Maintenance
In a study of the Uranium Information Centre (2006) in which the economic of the nuclear generation
is compared with other generation capacity the O&M costs for nuclear generation is set on 7.2/MWh,
which is more or less in line with the levelized cost comparison for new generating capacity in the US
by the EIA (2006) were O&M cost are set on $ 7.8/MWh (approximately 6.3/MWh) for nuclear
technology. According to White (2005) the O&M cost for nuclear are $ 8.7/MWh (approximately
7.0/MWh).
For the model calculations O&M cost of 6.8/MWh are used.
Unit Size
According to the Japan Nuclear Energy Safety Organization (2005) as of march 2006, four ABWRs
were in operation in Japan: Kashiwazaki-Kariwa units 6 and 7 (rated power 1356 MWe), which opened
in 1996 and 1997, Hamaoka unit 5 (rated power 1380 MWe), opened 2005 after having started
construction in only 2000 and Shika unit 2 (rated power 1358 MWe), which started commercial
operation march 2006 . Another two, identical to the Kashiwazaki-Kariwa reactors, were nearing
completion at Lungmen in Taiwan, and one more (Shimane 3, licensed output 1373 MWe ) had just
commenced construction in Japan, with major site works to start in 2008 and completion in 2011. On
planning stage are eight other ABWRs, with licenced capacity of 1373 - 1385 MWe and with (planned)
start of commercial operation between 2012 and 2017.
The unit size for the ABWR at TVA Bellefonte have been given in a study by Toshiba et al (2005) and
is set on a base output of 1371 MWe net and a uprate output of 1465 MWe net.
For the model calculations an ABWR unit of 1,465 MWe is used.

129

Part 3

Net Electrical Efficiency


According to Oka (2001) the net electrical efficiency is 34.5% but this is based on the ABWRs in
Japan currently in operation. It is assumed that with the ongoing developments the efficiency of the
ABWR will rise to 36 - 37% and for the ABWR II (generation III+ reactor design) even further.
For the model calculations a (base load) net electrical efficiency of 36% for a ABWR unit of 1,465
MWe is used.
Investment and Decommissioning Costs
The EPC costs for the ABWR at TVA Bellefonte have been calculated in a study by Toshiba et al
(2005). The costs in this study are indicative prices based on mutually agreed terms and conditions
and site conditions using fixed prices, see Table 6.2.

Plant capacity

Entire Plant

Power Block

Base Output (1,371 MWe-Net)

$ 1,611 /kW

$ 1,443 /kW

Uprate Output (1,465 MWe-Net)

$ 1,535 /kW

$ 1,377 /kW

Table 6.2: Bellefonte ABWR EPC overnight cost summary in 2004 dollars (Source: Toshiba
Corporation et al, 2005)
At an average dollar/euro exchange rate of 1.243 in 2004 the EPC costs for the Bellefonte ABWR
work out at 1,235/kWe. Including estimated non-EPC costs of 30% this results in a specific total
investment of approximately 1,600/kW for 2012. Learning curves for new nuclear power show that
total specific investment will decrease. Riahi et al (2004) calculated a decrease in investment cost for
nuclear power as a function of cumulative installed capacity. According to the EIA (2006), the worlds
nuclear powered generation capacity will increase (IEO2006 reference case) from 361 GWe in 2003 to
438 GWe in 2030. The decrease in investment costs over a period of 27 years would amount to 10 15%. Assuming a 10% reduction in investment costs by 2025 would lead to a specific total investment
of approximately 1,450/kW.
The Uranium Information Centre (2006) estimated decommissioning costs of a nuclear power plant at
9 - 15% of the initial capital cost. Discounted they contribute only a few percent to the investment cost
and even less to the generation cost, in the US$ 0.1 - 0.2/kWh which is no more than 5% of the cost of
the electricity produced.
European Pressurized Reactor (EPR)
The EPR is a generation III+ (Areva-NP, 2007) fission reactor design, based on the pressurized water
reactor or PWR. It has been designed and developed mainly by the Commissariat l'nergie
Atomique in France and the Karlsruhe Research Center in Germany. A French-German cooperation
was set up from the start of the project to develop the EPR. The nuclear activities were brought
together in AREVA NP (formerly Framatome ANP) of which AREVA has a 66% and Siemens a 34%
stake in the company.
The main design objectives of the EPR are increased safety while providing better electrical
performance. The reactor can use 5% enriched uranium oxide or mixed uranium plutonium oxide fuel.
The EPR design has several active and passive protection measures against accidents:
Four independent emergency cooling systems, each capable of cooling down the reactor after
shutdown.
Leak-tight containment around the reactor.
An extra containment and cooling area if a molten core manages to escape the reactor (core
catcher, see containment building).
Two-layer concrete wall with total thickness of 2.6 meters, designed to withstand collision with
airplanes.

130

Chapter 6 Technological Trends in Supply Options

The Olkiluoto power plant in Finland, scheduled to go on line in 2009, will be the first EPR reactor
built. The construction will be a joint effort of Areva NP, supplying the nuclear island (comprising the
reactor and associated systems), while Siemens is responsible for the conventional island (steam
turbine and generator).The electrical net power output of the plant will be 1600 MWe.
Subject to the result of Public Enquiry in 2006 and issue of a construction permit in 2007, a
demonstration EPR reactor, as a forerunner for a subsequent serie of EPRs, is built in France in
Flamanville in the Manche dpartement, to be operational in 2012.
Main Components and Processes
The EPR has four steam generators one for each of the four heat removal loops composing the
primary system. The steam generators (SG) are the interfaces between the primary water heated by
the nuclear fuel and the secondary water which provides steam to the steam turbine generator (there
is a leak-tight separation between the primary and secondary sides of the steam generators). That
secondary heat produces steam to power the turbine generator which in turn generates electricity.
Figure 6.4 shows a schematic of the EPR.

Figure 6.4: EPR schematic (Source: AREVA-NP)


The reactor containment building has two walls: an inner pre-stressed concrete housing (5) internally
covered with a metallic liner and an outer reinforced concrete shell (6), both 1.30 meters thick.
It houses the reactor coolant system, whose main components are the reactor vessel (1), the steam
generators (2), the pressurizer (3) and the reactor coolant pumps (4).
Inside the containment there is a special area (7) where in the event of core meltdown, any of the
molten core escaping from the reactor vessel would be collected, retained and cooled.
The turbine building (10) houses the equipment that transforms the steam produced into electricity: the
turbine, the alternator and the transformer, which is connected to the grid. In the event of a power
blackout, diesel generators, housed in two separate buildings (9), supply electricity to the safety
functions.

131

Part 3

Safety System
PWR reactors are extremely safe industrial facilities. The EPR offers even greater safety due to its
design, the materials used and improved safety systems. These systems have been simplified,
diversified and more fully automated. They also have a greater degree of redundancy then PWR
reactors.

a
b
Figure 6.5: Safety systems EPR (Source: AREVA-NP)

To provide emergency cooling of the reactor core, there are four independent sub-systems or "trains",
each of them capable of performing the entire safety function on its own, see Figure 6.5a. Each subsystem is separated from the others. They are located in separate parts of the plant, each with its own
protection features. The risk of simultaneous failure of all the sub-systems due to events of internal or
external origin (such as fire or aircraft crash) is thus avoided.
Despite the fact that these systems reduce the probability of a serious accident to near zero, the EPR
is designed in such a way that should an accident occur, the leak-tight containment around the reactor
prevents radioactivity and its effects from spreading outside (b). The containment can withstand high
pressure and temperature, even during extremely severe accidents involving core meltdown and
degradation or failure of the steel reactor vessel.
Under these circumstances, if any part of the molten core did manage to escape from the reactor
vessel, it would be passively collected and retained, then cooled in a specially designed area inside
the containment (c), the core cather. Even in such an extreme situation, the accident would be
confined within the reactor containment building.
The containment building (d) which houses the reactor is particularly robust. It stands on a single
reinforced six-meter thick concrete basemat. The upper part comprises two walls: an inner prestressed concrete housing and an outer reinforced concrete shell (covering the reactor building, the
spent fuel building, the control room, two of the four safeguards buildings), both 1.3 meters thick,
giving a total of 2.6 meters of concrete.
Operation and Maintenance
See section on ABWR.
Unit Size
The net electrical power output of the Olkiluoto power plant in Finland will be 1,600 MWe and this is
also used for the model calculations.
Net Electrical Efficiency
According to AREVA-NP the net electrical efficiency is 36 - 37% depending on site conditions
(presently the highest value ever for water reactors).
For the model calculations a (base load) net electrical efficiency of 36% is used.
Investment and Decommissioning Costs
According to a press release of AREVA (2007) the overall Olkiluoto 3 project cost has been indicated
by the Finnish utility Teollisuuden Voima Oy (TVO) at around 3 billion. These are probably EPC +
other cost and this leads with a net electrical power output of 1,600 MWe to a specific investment of
approximately 1,875/kWe for 2012. Assuming a 10% reduction in investment costs by 2025,
according to the description of investement drop at the ABWR would lead to a specific total investment
of approximately 1,650/kWe by 2025 (for decommissioning costs see section on ABWR).

132

Chapter 6 Technological Trends in Supply Options

Pebble Bed Modular Reactor PBMR


The PBMR based on a single-loop direct Brayton/Joule thermodynamic cycle with a helium-cooled
and graphite-moderated nuclear core assembly as a heat source is a High Temperature Reactor
(HTR), with a closed-cycle gas turbine power conversion system. According to the South African firm
Pebble Bed Modular Reactor (Pty) Ltd it is not the only HTR currently being developed in the world,
but the South African project is internationally regarded as the leader in the power generation field.
Main Components and Processes
The pebble bed design was first conceived in the 1950s by Rudolf Schulten, a physicist in Germany.
His idea was to create a safer nuclear reactor by the packing of tiny particles of uranium into thousand
of graphite spheres to be used as fuel for a new type of helium cooled type of reactor. The technology
was developed in Germany in the 1960s through 1980s and demonstrated in two reactors over this
period, the AVR and the THTR. However, continuing technology development ceased in Germany for
political reasons shortly after the Chernobyl reactor accident. The South Africans picked up this
technology in the late 1990s. The PBMR essentially comprises a steel pressure vessel which holds
about 450,000 fuel spheres. The fuel consists of low enriched uranium triple-coated isotropic particles
contained in a moulded graphite sphere. A coated particle consists of a kernel of uranium dioxide
surrounded by four coating layers. The PBMR system is cooled with helium. The heat that is
transferred by the helium to the power conversion system, is converted into electricity through a
turbine.
The plant comprises (1) a module building with the reactor pressure vessel (RPV) and the power
conversion unit (PCU), see Figure 6.6 (Pebble Bed Modular Reactor (Pty) Ltd, 2006).

Figure 6.6: Main components PBMR (Source: Pebble Bed Modular Reactor (Pty) Ltd)
Process
To remove the heat generated by the nuclear reaction, helium coolant enters the reactor vessel at a
temperature of about 500C and a pressure of 9 MPa. The gas moves down between the hot fuel
spheres, after which it leaves the bottom of the vessel having been heated to a temperature of about
900C. The hot gas then enters the turbine which is mechanically connected to the generator through
a speed-reduction gearbox on one side and the gas compressors on the other side. The coolant
leaves the turbine at about 500C and 2.6 MPa, after which it is cooled, recompressed, reheated and
returned to the reactor vessel.

133

Part 3

Inherent Safety
With the PBMR this basic danger of overheating is independent of the state of the reactor coolant. The
PBMR combines very low power density of the core (1/30th of the power density of a Pressurized
Water Reactor), and the resistance to high temperature of fuel in billions of independent particles
which creates an inherent celing to temperature control.
The helium, which is used to transfer heat from the core to the power-generating gas turbines, is
chemically inert. It cannot combine with other chemicals and is non-combustible. Since air cannot
enter the primary circuit, oxygen cannot get into the high temperature core to corrode the graphite
used in the reactor. Thus chemical reactions and oxidation, two of the great dangers in conventional
reactors are sidelined by the construction of the PBMR.
Operation and Maintenance
According Boer (2004) the O&M costs for a single reactor plant are an estimated 5/MWh and a study
of Kadak (1998) refers to levelized O&M costs of $ 4.4/MWh for an 8-module plant.
For the model calculations O&M cost of 5/MWh are used.
Unit Size
The reactor thermal power of a PBMR is in the order of 268 MWth with optimizations leading to 400
MWth (Matzner, 2004; Boer, 2004). With 400 MWth this results in an electrical power generation of
170 MWe with possible co-generation of 180 MWth (Boer, 2004). The MIT PBMR has a power output
of 120 MWe (Kadak, 2005). According to Matzner the reactor power output can increase to 600 MWth
in the future leading to an electrical power generation of 300 MWe.
For the model calculations a PBMR unit of 170 MWe is used for 2012 increasing to 300 MWe for
2025.
Net Electrical Efficiency
According to Boer (2004) the net electrical efficiency of a PBMR is 42.5% and this is in line with
Kikstra (2001) where this efficiency is approximately 42%. The target efficiency of the MIT Pebble Bed
Reactor is 45% (Kadak, 2005). For part load conditions no further information is found but the study of
Boer into the feasibility of the HTR reactor as a backup system for wind energy concludes that the
PBMR reactor can adjust its power output in a fast manner and suitable to follow variations in wind
production. This is stated by the Australian Uranium Association Nuclear Issues Briefing Paper 16 of
December 2005 which says: Performance includes great flexibility in loads (40 - 100%), with rapid
change in power settings. According to Boer also the cost of energy generation of a PBMR in
combination with wind energy compared with a conventional gas-fired unit is lower. Therefore it is
assumed that the efficiency curve of a PBMR has to be rather flat.
For the model calculations a (base load) net electrical efficiency of 42% is used for 2012 and 45% for
2025.
Investment and Decommissioning Costs
According to Nuclear Issues Briefing Paper 16 the overnight construction cost (when in clusters of
eight units) is expected to be $ 1,000/kW and generating cost below $ 0.03/kWh. Investors in the
PBMR project are Eskom, the South African Industrial Development Corporation, and Westinghouse.
Boer (2004) refers to capital cost according to the Pebble Bed Modular Reactor (Pty), the total capital
cost is between $1,000 and $1,200 per kWe, based on the construction of an 8-module power plant
but excluding the costs for decommissioning the plant. A demonstration plant is due to be built in 2006
for commercial operation in 2010. Analysis of Boer leads to a cost of $ 1,500/kWe. Kadak (2005)
estimates per 110 MWe unit module a capital cost of about $ 200 million, almost the double amount,
$ 1,820 /kWe but this looks conservitive. With an estimated capital cost of $ 1,350/kWe, non-EPC cost
of 30% and an average dollar/euro exchange rate of 1.243 in 2004 this leads to a specific total
investment of approximately 1,400/kW. Assuming a 10% reduction in investment costs by 2025,
according to the description of investement drop at the ABWR would lead to a specific total investment
of approximately 1,250/kWe by 2025.
Calculations by Boer reveal decommissioning costs of 0.4/MWh for the PBMR, about a factor 3.5
more than for the ABWR and EPR. Economies of scale play a role here.

134

Chapter 6 Technological Trends in Supply Options

Nuclear Fusion
Process
In the nuclear fusion process for fusion reactors a deuterium nucleus (hydrogen with 1 neutron) and a
tritium nucleus (hydrogen with 2 neutrons) are converted into a helium nucleus, energy, and one
neutron. Massive kinetic force is needed to fuse the nuclei, which repel each other as both are
positively charged. The nuclei have enough speed at around 150 million degrees Celsius. At this
temperature the atoms are a plasma; in other words, atomic nuclei and electrons fly around. To make
the atoms hot enough to form a plasma an electric current is passed through the gas or the atom
nuclei are radiated with micro-waves.
There is no known material that is capable of withstanding the temperatures required for the fusion
process. As the plasma needs to be at a safe distance from the reactor wall at all times, it is kept in a
magnetic field: atomic nuclei are positively charged and the Lorentz force exerted on the nuclei makes
the plasma move in a more or less circular orbit. The magnetic field is built in such a way that the
Lorentz force pushes back any nuclei that try to escape from the circle.
Scientific Research
The aim of international research on nuclear fusion is to realize a prototype for a nuclear fusion plant
that meets the demands of society in terms of safety, reliability, fuel supply, pollution and costeffectiveness. Considerable scientific and technological strides have been made in fusion research in
the past decade. The main obstacle is still how to confine the plasma. The implementation of the
magnetic field is a formidable challenge.
Since 1988 the nuclear fusion community has been collaborating on a large international experiment
called ITER which stands for International Tokamak Experimental Reactor. ITER is a project between
Europe, Russia, the USA, Japan, China and South Korea. It is expected to go into operation around
2016 and should demonstrate the technology needed for commercial fusion power production. This
plant is designed to produce 500 MW of thermal power which is ten times the energy needed to
sustain the plasma in which the fusion reaction takes place. The reactor will be built in Cadarache in
the south of France.
Future Power Plant
The fuel for nuclear fusion is inexhaustible and the process is inherently safe. It also produces less
radio-active waste than nuclear fission and this is short-lived waste. The first commercial power plant
is not expected before 2050 (see Figure 6.7 for a schematic of a nuclear fusion power plant).

Figure 6.7: Schematic of a nuclear fusion plant (Source: TU Delft)

135

Part 3

6.2.3

Integrated Gasification Combined Cycle (IGCC)

Like the PFBC (Section 6.2.1), the technology is relatively new in connection with power generation.
Coal-based Integrated Gasification Combined Cycle (IGCC) plants for power generation passed
through a critical stage in their development during the 1990s. Using IGCC, typically 60 - 70% of the
generated power comes from the gas turbine.
Gasification processes are commercial technologies with over 50 years of commercial application. The
Texaco process (Seabright, 2001) can gasify coal, heavy oil, petroleum coke and refinery residues.
Shell Global Solutions licensed two gasification technologies (Zuideveld, 2003): the Shell Gasification
Process (SGP) for liquid feedstock applications, and the Shell Coal Gasification process (SCGP) for
solids such as coal and petroleum coke, the latter is applied for the Willem Alexander Power station
(see also Section 4.4.6).
Three gasifier formats are possible: with fixed beds (not normally used for power generation), fluidized
beds and entrained flow. Fixed bed units use only lump coal, fluidized bed units a feed of 3 - 6 mm in
size, and entrained flow gasifiers use a pulverized feed, similar to that used in PCC.
IGCC plants can be configured to facilitate CO2 capture. The new gas is quenched and cleaned. The
syngas is shifted using steam to convert CO to CO2 and H2 of which the CO2 is then separated for
possible long-term sequestration.
Process Characteristics
The applied cycle for IGCC are the Brayton/Joule cycle as topping cycle and the Rankine cycle as
bottoming cycle. The Rankine cycle is implemented as an unfired heat recovery steam generator
(HRSG) behind a Joule cycle, the gas turbine. Future developments of the Brayton/Joule and Rankine
are described by CCGT technology in Section 6.2.4. The significant technical and future challenges for
IGCC are for the gasification process and the rather complex integration of the gasification process
and the power plant. Highly integrated plants tend to have long start-up times (compared to PCC
units), and hence may only be suitable for base-load operation.
The principal future developments for IGCC units involve improvement of the gasification process.
With pressurized gasification (as with PFBC), the supply of coal into the system is considerably more
complex than with PCC. Some gasifiers use bulky and costly lock hopper systems to inject the coal,
while others have the coal fed in as a water-based slurry. Similarly, by-product streams have to be
depressurized, while heat exchangers and gas cleaning units for the intermediate product syngas
must themselves be pressurized.
Some of the major process variations include (IEA Clean Coal Centre):
The possibility of air separation into oxygen and nitrogen streams prior to the gasification unit.
Entrained flow units, with a pulverized coal feed, or fluidized bed units with a coarser coal feed
and lower operating temperatures (of about 900 versus 1600C).
Heat can be recovered from various parts of the system, and used to heat and superheat the
steam to be expanded through the steam turbine.
The syngas is normally cooled to around 50C in current demonstration units, so that it can be
cleaned before being burned and fed to the gas turbine. A better alternative is to treat the syngas
in a hot gas cleanup device. These are currently being tested at temperatures of around 500 600C, so some cooling is still required.
Historically, most gasifiers have been oxygen blown, because of the costs of handling large amounts
of nitrogen, and the effect it has on diluting the product syngas. Japanese development, however, is
concentrating on air blown systems.
The fundamental advantages of oxygen blown gasification are:

136

Reduced gasifier size, and hence cost.


The heating value of the cooled and purified syngas is higher.
The syngas volume is about half that for an air blown unit for the same amount of coal gasification
energy, thus gas handling and cleanup requires smaller units.

Chapter 6 Technological Trends in Supply Options

Smaller heat exchangers are required to recover as much of the sensible heat from the syngas as
possible before cleanup.

The disadvantage of oxygen blowing is that the degree of plant integration required is considerably
increased. This means that controlling and operating the plant is more like running a complex
chemical plant than a traditional power station. Matching the requirements for availability, reliability
and flexibility of operation (for example, to load follow) at a competitive cost over a long period is the
major challenge. Auxiliary power consumption in an air blown system is estimated to be less than 8%,
compared with 10 - 15% for oxygen blown systems. Development in Japan on a 200 t/d pilot scale has
been based on the air blown route, and a design has been developed for a commercial size
demonstration unit to use 2,000 t/d of coal.
The syngas is produced at temperatures up to 1700C (in entrained flow gasifiers), while the gas clean
up systems which are being assessed, operate at a maximum temperature of 600C. Large heat
exchangers are required, and there is the possibility of solids deposition in these exchangers which
reduces heat transfer. It seems that unless it is possible to develop hot gas cleaning as a reliable
procedure, the comparative economics of IGCC will remain unattractive.
Gasifiers may be able to use coals that would otherwise be difficult to use in PCC plant, such as those
with a high sulfur content, or high ash content. The current demonstration units will test various coals,
and should resolve many of the technical issues outlined above.
Ash behavior in a gasifier is a critical parameter, both in terms of the satisfactory formation of a slag in
entrained flow, and the possibility of solids deposition in the syngas cooler/heat exchanger. At lower
temperatures, such as those in fluidized and fixed bed gasifiers, tar formation and deposition may
prove to be a difficulty. One advantage of gasification under pressure is that the effective gas volumes
involved are far smaller from gasification than from PCC.
Operation and Maintenance
From a maintenance perspective an IGCC consists of a gasification part and a power part. The
operation of the gasification part is integrated with that of the power part and is controlled centrally.
The gasification installation requires specific maintenance and the maintenance for the power part is
similar to that for a CCGT installation (see also Section 6.2.4). In VGB report (2004) O&M cost for a
776 MW IGCC unit is 12/MWh in 2003 and decreasing to 10/MWh in 2020. From the combined
ECN-Kema study mentioned before (ECN,1995) one might derive the expected differences between
PCC and IGCC to get an idea of the added costs. It appears that the fixed maintenance costs for an
IGCC will be an estimated 40% higher by 2010 and around 10% higher by 2030. The variable
maintenance costs, on the other hand, will be similar to those for a PCC in 2010 and approximately
30% lower in 2030.
For the model calculations a fixed cost of 7.0/MWh and variable cost of 2.0/MWh is used for 2012
and a fixed cost of 5.5/MWh and variable cost of 1.5/MWh is used for 2025.
Unit Size
A number of demonstration units, mainly around 250 MWe size are being operated in Europe and the
USA. Most use entrained flow and are oxygen blown, and one is based on a fluidized bed, and is airblown. The 253 MWe unit at Buggenum (see Section 4.4.6) in the Netherlands, started up in 1993.
Three plants are in the USA at Wabash River in Indiana; Polk Power near Tampa in Florida; and
Pion Pine in Nevada. One unit is at Puertollano in Spain with a capacity of 330 MWe. Adjacent to
Italys second largest refinery the 512 MW IGCC power plant in Priolo, Sicily is operating. The unit
consist of 2 blocks with in each block a Siemens V94.2 gas turbine unit of 156 MWe and a steam
turbine of 114 MW . According to GE (GER-3936A, 2001) a capacity rating of their new H technology
a single shaft IGCC unit of 500 - 530 MWe is possible. For the model calculations 550 MW and 900
MW units (2 blocks) are used for 2012 increasing to 600 MW and 1200 MW (2 blocks) units for 2025.

137

Part 3

Net Electrical Efficiency (based on LHV)


As with PCC, PFBC, the driving force behind the development is to achieve high net electrical
efficiencies together with low levels of emissions. With all power generation routes, it is important to
assess and compare efficiencies under normal load following conditions, and not just when the unit is
operating under full load. It is hoped to reach efficiencies of over 40%, and possibly as high as 45%
with IGCC. Higher efficiencies are possible when high gas inlet temperatures to the gas turbine can be
achieved. At the moment, the gas cleaning stages for particulates and sulfur removal can only be
carried out at relatively low temperatures, which restricts the overall efficiency obtainable.
The main incentive for IGCC development has been that units may be able to achieve higher thermal
efficiencies than PCC plant, and be able to match the environmental performance of gas-fired plants.
During the development phase, the thermal efficiencies of new PCC plants using superheated steam
have also increased. According to Siemens (Voges, 2004) the present IGCC plants attain efficiency
levels of approximately 45 to 48%, according to GE (GER 3936A, 2001) this range is 45 - 49%, and
Siemens pronounce they might go as high as 54 to 57% (2020).
For the model calculations a (base load) net electrical efficiency of 46% (900 MWe) and 47% (550
MWe) is used for 2012 and 55% (1200 MWe) and 56% (600 MWe) for 2025.
Investment Cost
According to GE (GER 4207, 2000) improvements have been made in IGCC capital cost. Solid fuel
plants have been recently bid for less than $1,000/kW on a turnkey basis and this is 30 - 40% of the
cost of the first IGCC plants. Gas turbine performance enhancements, gasification system
enhancements and EPC learning curve effects are driving cost reduction. According to DOEs National
Energy Laboratory (NETL) at present a new IGCC plant in the US would cost approximately $1,200$1,400/kW and the goals for 2010 are to achieve cost of $1,000/kW. In VGB Report (2004) a specific
investment for a 776 MW IGCC unit is given of 1,370/kWe for 2003 decreasing to 1,130/kWe for a
879 MWe IGCC unit in 2020. An advanced coal-fired plant with the required pollution controls would
cost about $1,000-$1,200/kW (Parkinson, 2004). IGCC will cost a factor 1.15 more than coal-fired
power plants. Blankinship (2006) says that it is generally assumed that an IGCC plant will cost 20 to
25 percent more than a comparable supercritical unit, with a targeted capital cost premium of around
10 percent above that of a comparable supercritical unit. The investment cost for a PCC are here
assumed to be between $1,200 and $1,460/kWe.
For the model calculations, with an investment cost of 1,200/kWe for PCC by 2012 (sesection on
PCC), the investment cost for an IGCC are 1,500/kWe decreasing to ,1200/kWe in 2025 against
1,100/kWe investment cost for a PCC.
6.2.4

Combined Cycle Gas Turbine (CCGT)

Gas turbine technology (Brayton/Joule cycle) has progressed by leaps and bounds in recent decades.
This, in combination with a steam cycle (Rankine cycle), has triggered a spectacular rise in the total
efficiency of combined cycles. There is still potential for raising efficiency even further.
Gas Turbine Technology
Process Characteristics
The first gas turbine to generate electricity was commissioned in Neufchateau in Switzerland in 1939.
It was deployed as a peak load turbine because it had a short start-up time, the investment costs were
low, and the delivery time was short. The low thermal efficiency compared with the steam power
plants at that time was accepted because the operational hours were low. Since then, developments
in gas turbine technology particularly in blade-cooling and the materials (alloys) for the blades and
in combustion technology have moved forward and are improving the power and efficiency of gas
turbines (see Table 6.3).

138

Chapter 6 Technological Trends in Supply Options

Name

First gas

Current

Current

Current

Simple

Simple

turbine

largest

largest gas

largest gas

Cycle Aero-

Cycle Aero-

Simple

turbine

turbine

gas turbine

Frame gas

Cycle gas

(with re-

(with steam

with

turbine with

turbine

heating)

cooling)

highest

highest

output

output

SGT5-

GT26

9H

LM6000PD

LMS100

98

4000F
Power [MW]

268

288

320

42

Efficiency [%]

17.4

38.7

36.9

NK

40.8

45.7

Exhaust gas flow [kg/s]

62

665

445

685

127

207

Exhaust gas temperature [C]

360

585

612

NK

448

417

Pressure ratio [-]

4.2

NK

32

23

29.3

42

Turbine inlet temperature [C]

550

NK

1260

1430

1300

1380

Specific Power [kJ/kg]

64.5

NK

438

NK

328

NK

Table 6.3: The first industrial gas turbine compared with the modern gas turbines (Sources: Van
Buijtenen, 1997; GER 3935B, 2000; Siemens, Alstom and General Electric power generation)
The table shows substantial increases in specific power and turbine inlet temperature as well as in
power and efficiency. It should be noted that the largest gas turbines, the SGT5-4000F and GT26 in
Table 6.3 are (situation 2006) the largest gas turbines that can operate in simple-cycle mode, i.e.
without a steam cycle. Models like the GE MS9001H (also in Table 6.3) and the Siemens SGT58000H are even more powerful (approximately 320 MW) but are not deployed in simple cycle mode.
Because it is fitted with innovative steam-blade cooling (see below) the GE MS9001H probably cannot
operate without attached steam cycles.
The table also shows the types of gas turbine that are currently used to generate electricity:
The aeroderivatives are gas turbines which originate from aircraft engines (see Figure 6.8).
They have an optimized pressure ratio for maximum efficiency. The gas turbine consists of,
amongst other things, various compressor and turbine pressure sections (spools) which rotate
in opposite directions at different rpms. The sections with the highest pressure also have the
highest rpms (LM6000 is a 2-spool gas turbine with 2 compressor and 2 turbine spools). These
turbines are also lightweight with small dimensions and a lower exhaust temperature (Column
5 in Table 6.3). They are also easily interchangeable for maintenance.

Figure 6.8: LM6000 PD aeroderivative (Source: GE Power)

139

Part 3

The combination of aeroderivatives and frame gas turbine technology e.g. the LMS100 by
General Electric (GE, GER-4222A, 2004). The new 3-spool (LPC, HPC compressor and HPT,
IPT and PT) LMS100 gas turbine (see Figure 6.9) delivers the highest simple cycle efficiency
in the industry today (Column 6 in Table 6.3).

Figure 6.9: LMS100 Gas Turbine (Source: GE, GER-4222A, 2004)

The larger industrial gas turbines (see Figure 6.10) are usually optimized for maximum
gas/steam cycle efficiency by means of a linked water/steam cycle, which creates a higher
exhaust temperature and steam mass flow (see Table, Columns 2, 3 and 4). These gas
turbines are heavyweight with large dimensions and a higher exhaust temperature. They are
not, in principle, interchangeable for maintenance.

Figure 6.10: MS9001H rotor on the half shell (Source: Power, July/August 2003)

140

Chapter 6 Technological Trends in Supply Options

The performance of the gas turbine will be further enhanced via the turbine inlet temperature. This is
also where the limitations, as the combustion chamber components and the first stage in the
expansion turbine need to be able to bear the turbine inlet temperature, whilst at the same time the
first stage of the rotor blades must resist the tension from the centrifugal force. Developments in hightemperature materials technology and cooling techniques have kept pushing up the inlet temperature
(see Figure 6.11).

Figure 6.11: Development of turbine inlet temperature and others over time (Source: Ladwig, 2005)
Current developments geared at improving efficiency by raising the turbine inlet temperature:
Better materials such as high-grade alloys or special blade coatings.
Steam for cooling the gas turbine blades in combination with a water-steam cycle (the G and
H technology gas turbines by General Electric).
Current developments to raise efficiency:
Reheat the combustion gases (applied by Alstom in the GT26, see Column 3, Table 6.3).
The Intercooled Aero-Derivative (ICAD) with intermediary cooling in the compressor, heat
recuperation in the exhaust (being developed by Rolls Royce amongst others). Built on the
basis of the largest available aircraft engines, these systems will generate power in excess of
100 MW with an efficiency of approximately 60%.
Fuel cell integration for achieving efficiencies of more than 70%.
Research is also being conducted into gas turbines with water injection in the compressor, gas
turbines supercharged with Rotational Inlet Air Filter (an evaporative cooler), gas turbines with wet
compression, Humid Air Turbine cycles, semi-closed gas turbines, chemically-recuperated gas
turbines, CO2 gas turbines and more (for a detailed description of this research see Korobitsyn (1998)
and other publications).
Fuel and Emissions
The gas turbine owes its name to the fact that the working medium (air) appears in the gas phase
(Joule or Brayton process), it has nothing to do with the type of fuel it consumes. It can run on more or
less all types of liquid and gas fuel. The gases do however need to be far more refined than for
furnace combustion because they are in direct contact with the rotating expanders.
Special situations present themselves if the heating value of the fuel differs significantly from the
heating value for which the gas turbine was intended. For example, if the heating value of coal gas or
blast furnace gas (IGCC) is approximately 5 -15% of that of natural gas, the fuel flow needs to be 6-20

141

Part 3

times the normal flow for natural gas. This requires a differently designed gas turbine combustor with
many extra mechanisms for the air balance in the compressor. These technologies are gaining in
importance as a result of increasingly stringent rules on emissions and the clamp-down on primary
fuels in larger industrial processes.
When air is used as a working medium, the air excess is a factor of 4 or 5 and can be used to
minimize the formation of nitrogen oxides. This is achieved by, for example, splitting the combustion
process into stages or allowing it to take place in a weak mix of fuel-air (as thermal NOx can be
minimized by avoiding local stoichiometric conditions). The combustion of such weak mixes takes
place on the very edge of stability and therefore needs specially designed combustors, often referred
to as Dry Low Nox (industrial gas turbines) or Dry Low Emission (aero-derivatives), which were first
developed for natural gas. At present, single digits (NOx emissions < 10 ppmvd) are attainable with
these systems (see Figure 6.11). Other anti-emission technologies are based on lowering the
temperature of the flames by steam cooling or water injection. These technologies are applied
primarily in the case of low-calorie combustion gases or liquid fuel such as oil (although strides in
combustor design are being made in this field as well).
Combined Cycles
Process Characteristics
The Joule/Rankine combined cycle is by far the most advanced of all the combined cycles and is
accepted and applied everywhere. Thanks to its high exhaust temperature the topping Brayton/Joule
cycle is a perfect energy source for the bottoming Rankine cycle, where the working media (water and
air) are readily available, cheap and non-toxic. Other benefits of these combined cycle systems (GE,
GER- 3936A, 2001) include higher efficiency than conventional plants, low installation costs, fuel
flexibility, flexible work cycle (wide part- and base load regime and short start-up and stop times), and
a shorter installation time than conventional plants because subcomponents such as gas turbines and
steam turbines are prefabricated.
There are different types of combined cycles based on the Joule/Rankine cycle (see Table 6.4).
Brayton/Joule topping cycle
direct - fired

indirect-fired

Rankine bottoming cycle

HRSG
unfired
with supplementary firing
Steam boiler (repowering)
gas-, oil-, coal-fired
waste fired
Integrated with gas turbine
steam and water injection
blade cooling
combustor cooling

Table 6.4: Possible Joule/Rankine combinations (Korobitsyn, 1998)


The usual name for an HRSG (with direct or indirect firing) which has been combined with a steam
turbine and gas turbine is a CCGT. The hot windbox combined cycle units in current use in the
Netherlands (see Section 4.4.3) are re-powering units in which the overall efficiency is enhanced by a
gas turbine which serves as a warm air ventilator for the steam boiler.

142

Chapter 6 Technological Trends in Supply Options

Steam and water injections are used for suppressing NOx on the one hand and for generating extra
power on the other. The very latest technology GEs H technology uses steam for cooling the
blades and the combustor (see Figure 6.12).

Figure 6.12: CCGT 109H process cycle diagram (Source: GE, GER-3936A, 2001)
Further advances in gas turbine materials and hot path gas cooling technology as well as in higher
temperature and pressure steam cycles will continue the trend towards more efficient combined cycles
in the future (GE, GER- 3936A, 2001).
Operation and Maintenance
The gas turbine accounts for at least 60% of the total O&M costs of a CCGT plant (GE GER 4208,
2001). The introduction of more sophisticated technology (e.g. raising the turbine inlet temperature)
will increase the importance of gas-turbine-related maintenance and will account for an ever-larger
proportion of the total costs of generating electricity. GE says that its maintenance recommendations
are based on a detailed analysis of each turbine section and the factors that determine the metallurgy
and the mechanical reliability. The analysis is performed by monitoring the independent parameters
for fired hours, number of starts and combustion systems, and turbine hot gas components on the
basis of the past performances of some 5,000 GE technology gas turbines. GE claims that this
maintenance regime involves fewer inspections compared with the Equivalent Operating Hours (EOH)
regime. In VGB report (2004) O&M cost for a 400 MW PCC unit is 2.5/MWh. In a study of the
Uranium Information Centre (2006) in which the economic of the nuclear generation is compared with
natural gas fired generation and other generation capacity, O&M costs for natural gas fired generation
is set on 3.5/MWh which is not far short of the $ 4.2 per MWh which was estimated for O&M in the
overall costs of electricity generation in GEs F technology. ECN (2005) estimates 4.7 per MWh for a
large CCGT, but with heat generation. A 60% share for the gas turbine in this scenario works out at
around 2.0 per MWh. The ECN study in collaboration with Kema (ECN, 1995) only gives values for
the fixed cost and not for the variable costs for future CCGT production units, however based on own
information it is estimated that the variable part of the O&M cost amounts 40 - 45%.
For the model calculations used for 2012 and 2025, the total O&M cost are set on 3.5/MWh of which
the fixed cost part is 2.0/MWh and the variable cost part is 1.5/MWh.

143

Part 3

Unit Size
In the most recent General Electric H platform for gas turbine combined cycle technology, the gas
turbine will be configured with an integrated closed loop steam cooling system, see Figure 6.12, which
allows higher firing temperatures to be achieved without increasing combustion temperature.
According to GE a net power output of 480 MWe is delivered by the combination of a triple pressure
reheat steam cycle, a high pressure of 165 barg, and a steam temperature of 565C where the gas
turbine and steam turbine are connected as one drive train, driving the same generator (single shaft).
Siemens offers two types of configuration, the single shaft and the multi shaft where the gas turbine
and steam turbine are independent of each other, each driving their own generator. The single shaft
has a net power output of 416 MWe whilst the multi shaft doubles this to 832 MWe.
For the model calculations 480 MW and 832 MW units (2 blocks) are used for 2012 increasing to 500
MW and 1000 MW (2 blocks) units for 2025.
Net Electrical Efficiency (based on LHV)
GE (GER-4206, 2001) expects the integrated closed loop steam cooling system to improve net
electrical efficiency by 2 percentage points due, amongst others, to a reduction of the exhaust
temperature drop across the first stage nozzle and the elimination of chargeable cooling air for the
first and second stage rotating and stationary airfoils. The H platform gas turbine combined cycle
plants are expected to achieve 60% LHV thermal efficiency in the first half of this decade. According to
Siemens (2006) net electrical efficiencies for the 832 MWe multi shaft of 58.2% can be achieved and
this is confirmed by Voges (2004): CC power plants, which presently reach about 58% and were
challenged to come up with figures as high as 65% by 2020. Gasteiger of Alstom (2006) is expecting
target efficiencies of more than 63% in 2020.
For the model calculations a (base load) efficiency of 58% (832 MWe) and 60% (480 MWe) is used for
2012 and 63% (1,000 MWe) and 65% (500 MWe) for 2025.
Investment Cost
In VGB Report (2004) a specific investment for a 400 MW CCGT unit is given of 625/kWel and for a
790 MWe CCGT unit 410/kWe. In a CE study (2005) for a 1000 MWe CCGT unit an specific
investment of 500/kWe is applied. According to the consultation paper of the commission for energy
regulation (CER, 2002) the EPC cost for a 400 MW CCGT power station is 206 million and this leads
to an EPC price of 514/kWe. The total investment costs are also calculated and these are 271
million, leading to a total specific price of 679/kWe. Compared with the EPC cost this means that the
non-EPC costs are approximately 30% higher than the EPC price according to the CER.
According to power-technology.com (2006) the project cost of the 500 MW San Lorenzo CCGT power
plant, Batangas City in the Philippines is estimated to be $ 500 million. And this includes capital cost,
working capital requirements, related pipeline financing, insurance, and development. The plant is
generating electricity now and based on an average euro/US dollar exchange rate of 1.243 in 2004 the
total specific price is 804/kWe.
Centrica announced in June 2006 it is investing 400 million to develop the first major power station in
the UK, an 885 MW gas-fired plant in Langage, Devon and due to start during winter 2008/2009.
Based on a euro/pound stirling exchange rate of 0.6937 (5-7-06) the total specific price is 650/kWe.
Every year since 2004 the ECN has calculated the non-profitable top for CHP projects in its MEP
recommendations (2004, 2005). These calculations also include the specific investment costs for CHP
plants, based on consultations with market players. In 2006 a specific investment of 493/kWe is
assumed for a large 250 MWe CHP CCGT.
Though dependent on many factors, the investment costs for a larger CCGT will be in the 600800/kWe range including non-EPC cost. To compare the various options an average of 700/kWe by
2012 is assumed. Total specific investment will decrease according to learning curves for CCGT
plants. Riahi et al (2004) calculated a decrease in investment cost for CCGT as a function of
cumulative installed capacity. According to the EIA (2006), the worlds CCGT-powered generation
capacity increases by 1070 GW between 2003 and 2030. The decrease in investment cost over a
period of 27 years would amount to 10 - 15%. For investment cost in 2025 a 15% reduction is
assumed leading to specific total investment of approximately 600/kW. For the model calculations
the investment cost for a CCGT are set at 700/kWe in 2012 decreasing to 600/kWe in 2025.

144

Chapter 6 Technological Trends in Supply Options

6.2.5

Carbon Capture & Storage

Maintaining the benefits of access to low cost energy derived from fossil fuels, whilst at the same time
significantly decreasing CO2 emissions to the atmosphere, is a major challenge in the world today
(Source: Australian Coal Industry Coal21). The most promising technologies for significantly
decreasing emissions from large scale stationary sources of CO2 involve separating and capturing the
CO2, compressing and then storing it in geological or other locations where it will not leak back into the
atmosphere.
CO2 Capture
Technologies for capturing CO2 from emission streams have been used for many years to produce a
pure stream of CO2 from natural or industrial CO2 emissions for use in the food processing and
chemical industries. The gas industry routinely separates CO2 from natural gas before it is then
transported to market by pipeline. Methods currently used for CO2 separation include:

Physical and chemical solvents, particularly monoethanolamine (MEA)


Various types of membranes
Adsorption onto zeolites and other solids
Cryogenic separation.

These methods can be applied to a range of industrial processes. However, their use for separating
out CO2 from high volume-low CO2 concentration flue gases, such as those generated by conventional
pulverized coal-fired power stations, is much more problematic. The very high capital costs of
installing the huge post-combustion separation systems needed to process massive volumes of flue
gases is a major impediment to post-combustion capture of CO2. The second problem is the large
amount of additional energy (25 - 35%) used to release the CO2 from solvents or from solid
adsorbents after separation.
Capture technologies important for zero emission power plants can be divided in:

Pre-combustion; separation of CO2 from the fossil fuel


Post-combustion; removal of CO2 from combustion flue gases
Oxy-fuel combustion; combustion with pure O2 and recycled flue gas to reduce CO2 emissions.

Pre-Combustion; Separation of CO2 from the Fossil Fuel


In pre-combustion de-carbonization, the carbon content of the fuel is removed prior to combustion in
order to produce a hydrogen rich fuel and a CO2 by-product stream. The concept can be used for both
H2 production and electricity generation. In this study the capture technology is focussing on the
combination with integrated gasification combined cycle (IGCC). This involves, see Figure 6.13 first
the partial combustion of coal or gas in air (or oxygen) to produce a CO plus H2 gas stream, which is
reacted with hot steam to produce CO2 plus more H2. The H2 is then combusted in a gas turbine and
the CO2 is available for storage or use.

145

Part 3

Figure 6.13: IGCC with pre-combustion decarbonisation (VGB, 2004)


Post-Combustion; Removal of CO2 from Combustion Flue Gases
In post-combustion capture, see Figure 6.14, the CO2 is removed from the power plant flue gas.
Commercially available technology includes CO2 capture using absorption in an aqueous amine
solution. The CO2 is then stripped from the amine solution and dried, compressed and transported to
the storage site. Some major technical and cost challenges need to be addressed before retrofit (or
new build) of post-combustion capture systems becomes an effective mitigation option.

Figure 6.14: PCC with post-combustion decarbonisation, scheme shows SO2 and CO2 flue gas
scrubber and steam extraction for solvent regeneration (Source: VGB, 2004)
The present cost of post-combustion capture of CO2 is commonly quoted in the range of $ 30 - 60 a
ton of CO2. This would increase the wholesale price of electricity considerable. However, capture
technologies will undoubtedly improve in the coming years thereby improving economics.
A key to achieving lower capture costs lies in the production of a more concentrated, pressurized
stream of CO2 (the average PCC generator has only 10-14% CO2 in the flue gas stream). This can be
achieved through the pre-combustion capture of CO2 or oxy-fuel combustion.

146

Chapter 6 Technological Trends in Supply Options

Oxy-Fuel Combustion; Combustion with Pure O2 and Recycled Flue Gas to Reduce CO2 Emissions
An alternative approach for removing CO2 from the flue gas is to use oxygen for combustion instead of
air which contains about 79% by volume nitrogen, see Figure 6.15. Combustion the fuel using almost
pure oxygen (95 to 99%) at near- stoichiometric conditions, creates a flue gas consisting mainly of
CO2 (> 90 - 95% on a dry basis), water vapour and minor amounts of noble gases and, depending on
fuel composition, SOx and NOx. Whilst the principle is simple, there are major issues to overcome,
including the very high combustion temperatures and the cost of producing the oxygen. Oxy-fuel
combustion for power generation may be a future option though it has yet to confirm its operational
and commercial viability. Much the same oxy-combustion technique is used in steel making and
consequently there may be no insurmountable technical barriers to CO2 storage linked to oxy-fuel
power generation in the future. As long as the main pressure on power companies is to decrease
electricity costs, it is unrealistic to expect them to voluntarily retrofit post-combustion CO2 capture or
move to IGCC or oxy fuels. However, if in the future the main pressure is to reduce CO2 emissions, all
technical options, including perhaps post-combustion retrofit, will receive increased attention given
that many countries have made massive investments in conventional thermal power stations.

Figure 6.15: PCC with O2 /CO2 recycle combustion and decarbonisation (VGB, 2004)
RWE, among others, is examining a process which combusts coal in a mixture of pure oxygen and
recirculated flue gas in a joint Advanced Development of the Coal-fired Oxy-Fuel Process with CO2
Separation (ADECOS) project. In the experimental oxy-fuel process, fossil-fired power plants burn
hard coal or lignite along with pure oxygen and not, as is usual, with air. Afterwards, the flue gas that
emerges contains almost exclusively CO2 and steam. The components can be separated by cooling:
the steam condenses to water, the carbon dioxide is left over and could then be stored below
ground. So, the oxy-fuel process is a possible step toward a low-emission power plant, although it is
still early days for this development (RWE, 2006). Together with other project partners, the developers
at RWE are at the moment thoroughly testing this innovative combustion method. Pilot plants are
furnishing important findings, like the emission behavior of the oxy-fuel system or its proneness to
fouling and slagging. The oxy-fuel process, like other CO2-avoidance technologies as well, leads to a
loss in efficiency. Before larger demonstration power plants will be built, this is to be further optimized
and ADECOS will boost research into the oxy-fuel process going with new findings.
Net Electrical Efficiency (based on LHV)
The capture of CO2 always requires energy which results in a higher fuel consumption and
consequently a lower electric efficiency than the base line case without CO2 capture, see Figure 6.16
The efficiency of a rather new CCGT drops from approximately 56% to around 47%, whilst the PCC
efficiency drops from approximately 46% to about 33% and the IGCC drops from approximately 43%
to about 34% (VGB, 2004). According to these figures the efficiency drop of the coal-fired power plant
in percentage is almost twice the value of the CCGT. The IGCC shows less efficiency drop compared
with the PCC but more compared with the CCGT. The efficiency drop published by IPCC (2005) for
the same type of power plants are in line with the VGB results.

147

Part 3

60

58

58

56

49

49

50

47

46
43

43

58

56

48

48
46

43

Electric efficiency [%]

40
34

33

34

30

20

10

0
PCC, Post
Com bustion
Capture

IGCC, PreIGCC, PreCom bustion


Com bustion
Capture, F Gas Capture, H Gas
Turbine
Turbine

PF, lignite
Oxyfuel
Com bustion,
S-rem oval

CCGT, Post
Com bustion
Capture

CCGT, Post
Com bustion
Capture, NOx
Cleaning

CCGT, Pre
Com bustion,
Partial
Oxydation

CCGT, Pre
Com bustion,
NOX Cleaning

CCGT, Oxyfuel
Com bustion

Original
With CO2
capture

Figure 6.16: Electrical efficiencies of coal- and gas fired power plants with and without CO2 capture
and compression (Source: VGB, 2004 (based on studies of IEA, SINTEF))
Removal Efficiency
Like other emissions such as NOx and SO2, CO2 will not be removed for the full 100%. This would be
too costly (logarithmic cost curve!). Figure 6.17 shows that the CO2 removal for coal- and gas fired
power plants is approximately 80 - 87%. IPCC (2005) gives ranges for PCC between 81 - 88%, IGCC
between 81 - 91% and CCGT between 83 - 88% and this is in line with VGB.
1200

1000

400

200

CO2 captured

600

CO2 avoided

CO2 produced (kg/MWh)

800

Original

0
PCC, Post
Com bustion
Capture

IGCC, PreIGCC, PrePF, lignite


Com bustion
Com bustion
Oxyfuel
Capture, F Gas Capture, H Gas Com bustion,
Turbine
Turbine
S-rem oval

CCGT, Post
Com bustion
Capture

CCGT, Post
Com bustion
Capture, NOx
Cleaning

CCGT, Pre
Com bustion,
Partial
Oxydation

CCGT, Pre
CCGT, Oxyfuel
Com bustion, Com bustion
NOX Cleaning

Produced
before capture
After capture

Figure 6.17: CO2 emissions of coal- and gas fired power plants with and without capture (Source :
VGB, 2004 (based on studies of IEA, SINTEF))

148

Chapter 6 Technological Trends in Supply Options

Cost of reducing CO2 emissions


The CO2 capture technology is still experimental and has to be developed further to achieve better
performance of the power plant (i.e less energy consumption of the CO2 removal part of the plant and
therefore lower drop of electrical efficiency) and lower specific investment cost. When CO2 removal
becomes more and more important, due to stringent local and global CO2 reduction targets the
potentials are good. However, before that point is reached and removal of CO2 is economical
compared with todays (mid-2006) electricity and CO2 prices a lot of research has to be done. Riahi
et al (2004) fore see large scale applications of CCS as late as in the 2030s and once introduced
continuously gain market share. Entire diffusion of CCS form the initial introduction to saturation is
about 50 years, which is similar to diffusion speeds of other types of technologies like desulphurization
(FGD) technology. Learning curves result to a decreasing of cost of CCS, but are a function of
installed capacities with CCS in the world. Figure 6.18 shows a same kind of decrease of cost by an
increase of CO2 mitigation potential in the future. In 2020 only new or retrofit PCC is mentioned with
CCS whilst the IGCC technology with CCS in this figure comes in at 2030.

Figure 6.18: CO2 mitigation cost (Source: Damen, 2007 )


Investment cost
The cost of capturing CO2 is the largest component of overall CCS cost. In this study, capture cost
include the cost of compressing the CO2 to a pressure suitable for pipeline transport (typically about
14 MPa). The cost of CO2 transport and storage are not included (see next section). In Figure 6.19 the
specific investment cost are given for power plants with and without CO2 capture (VGB, 2004) . IPCC
(2005) gives represent able values for PCC of 2,096 /kWe with- and 1,286 /kWe without CO2
capture, for IGCC 1,825 /kWe with- and 1,326 /kWe without CO2 capture and for CCGT 998 /kWe
with- and 567 /kWe without CO2 capture. The percent increase in capital cost with capture amounts
for PCC 63%, IGCC 37% and CCGT 76%. Of the mentioned figure, only the IGCC of VGB (F-class
gas turbine) and IPCC are comparable. The percent increase of capital cost for PCC is in case of VGB
82% and IPCC 63%, for CCGT this is respectively 93% and 76%.

149

Part 3

2000
1860

1855

1790

1800

1600

1515
1430
1370

Specific investment [/kWe]

1400

1250

1270
1200

1130

1200
1020
1000

910
790

800

625

625

625

600
410

410

400

200

0
PCC, Post
IGCC, PreIGCC, PrePF, lignite
CCGT, Post
CCGT, Post
CCGT, Pre
CCGT, Pre
CCGT, Oxyfuel
Com bustion
Com bustion
Com bustion
Oxyfuel
Com bustion
Com bustion
Com bustion,
Com bustion,
Com bustion,
Capture, 500 -> Capture, F Gas Capture, H Gas Com bustion, S- Capture, 790 -> Capture, NOx
Partial
NOX Cleaning, 400 -> 400 MWe
362 MWe
Turbine, 776 -> Turbine, 879 -> rem oval, 865 ->
663 MWe
Cleaning, 400 -> Oxydation, 790 - 400 -> 392 MWe
Original
676 MWe
776 MWe
681 MWe
338 MWe
> 820 MWe
With CO2 capture

Figure 6.19: Specific investment of coal- and gas fired power plants with and without capture (Source:
VGB, 2004 (based on studies of IEA, SINTEF))
Based on the VGB and IPCC figures Table 6.5 is created for the PCC, IGCC and CCGT with- and
without CO2 capture as input for the model calculations. It is assumed for this study that still a lot of
research and development has to take place on CO2 capture technology to reduce the specific
investment cost and reduce the drop in electric power and electrical efficiency as given in the
nowadays studies. Large scale introduction of CO2 capture and therefore a large reduction of this cost,
power and efficiency because of learing effects is not foreseen within the next 20 years (Riahi et al,
2004; Lysen, 2007).

Power plant type

Original

With CO2
Capture

Original

With CO2
Capture

Original

PCC

PCC

IGCC

IGCC

CCGT

With CO2
Capture
CCGT

Plant size
Net efficiency (LHV)
CO2 emissions

MWe
%
kg/MWh

1200
53
638

875
38
113

1200
55
615

600
56
604

1050
43
100

525
44
98

1000
63
321

500
65
311

850
53
51

425
55
49

Captured CO2

kg/MWh

777

679

665

331

321

Avoided CO2
Specific investment
O&M cost

kg/MWh
/kWe
/MWh

1100
7

525
1900
13

515
506
1650
9.5

270
262
1100
7

1200
7

600
3.5

Table 6.5 : Data for the PCC, IGCC and CCGT with and without CO2 capture (Source: VGB, 2004,
IPCC 2005)
Transportation systems
After the CO2 is captured at the sources of emission, the CO2 has to be transported to the storage
sites. Such transportation will require a large scale infrastructure, due to the large volumes which are
in a minimum case 1 to 2 Mt/y for a 850 MWe CCGT plant and 5 Mt/y of CO2 for a 875 MWe coal fired
power plant. Transportation of CO2 can be performed on- and offshore by several means of

150

Chapter 6 Technological Trends in Supply Options

transportation - motor carriers, railway, shipping and pipelines. The most common and usually the
most economical method to transport large amounts of CO2 is through pipelines. For a possible future
scenario in the Netherlands for transportation of CO2 with storage facilities in empty gas fields on and
off shore and aquifers, see Figure 6.20.

Figure 6.20 Future scenario for transportation of CO2 with storage facilities (Source: Damen et al,
2006)
In Figure 6.21 the CO2 transport cost range for onshore and offshore pipelines per 250 km of normal
terrain conditions is given. The figure shows the low (solid lines) and high ranges (dotted lines).

Figure 6.21 : The CO2 transport cost range (Source: IPCC, 2005)
151

Part 3

CO2 Storage
A number of CO2 storage techniques are short considered : geological storage, ocean storage (see
Figure 6.22), and storing CO2 as mineral carbonate. For the advantages and disadvantages of each
technology and approximate cost ranges please refer to VGB, IPCC or other relevant literature.
.

Figure 6.22: Schematic diagram of possible geological CO2 storage options (Source: VGB, 2004)
Geological Storage
The most comprehensively studied storage option is geological storage. CO2 can be stored in
geological formations in several ways as a fluid within porous rock, by absorption into interstitial fluid
or within a fixed matrix , i.e. as a mineral carbonate. Three basic types of geological formations are
widespread and have adequate CO2 storage potential, see Table 6.6: deep aquifers; oil and gas
reservoirs and deep, unmineable coal seams.

Table 6.6 Estimated European underground CO2 storage capacity (Source: VGB, 2004)
Ocean Storage
Ocean storage of CO2 involves two main options: one is the dispersal of CO2 as droplets at
intermediate water depths of around 500 1,000m; the other is disposal at abyssal depths (5,000m or
more) as liquid CO2. Given the ocean is an enormous sink for CO2 and the system is strongly
buffered, the injected CO2 would probably have a negligible effect on the chemistry of the ocean as a
whole. Yet it would result in a measurable drop in the pH of seawater in the immediate vicinity of the
injection site and affect marine organisms. The ocean is an open system and it would be difficult, if not
impossible, to monitor the distribution of the stored carbon to confirm residence times of CO2. Also, the

152

Chapter 6 Technological Trends in Supply Options

impact of elevated levels of CO2 on marine ecosystems is poorly known and difficult to monitor. The
potential application of the London Dumping Convention to ocean storage of CO2 also raises legal
uncertainties. For all these reasons, there is widespread opposition to ocean storage and it is most
unlikely to be a CO2 storage option in the foreseeable future.
Storing CO2 as a mineral carbonate
Mineral carbonation is based on the reaction of CO2 with metal oxide bearing materials to form
insoluble carbonates, with calcium and magnesium being the most attractive materials. In nature such
a reaction is called silicate weathering and takes place on a geological time scale. The carbonation
process for captured CO2 is the so called mineral carbonation, where high concentration CO2 is
brought into contact with metal oxide bearing materials with the purpose of fixing the CO2 as
carbonates. The process is at an early stage of development and has not reach a level where a
thorough assessment of the technology, potential cost ans impacts is possible (IPCC, 2005).
6.3

Combined Heat and Power

Combined Heat and Power are combined cycle units with the potential to supply heat (on a large or
medium scale) in the form of industrial heat or district heating. These units, with the exception of gas
engines and micro-CHP, are described in detail in Section 6.2.4 In this section only the deviations for
Combined Heat and Power in relation with combined cycle units are described as well as the for the
model required input numbers. The characteristics of gas engines and micro-CHP are described
briefly and for the model required input numbers are derived.
6.3.1

Combined Heat and Power Units for Industrial and District Heating

The most important large and medium-scale CHP system by far is CCGT technology, followed by GT
+ HRSG technology, see also Section 4.5.3. Because of its very minor contribution to power
generation no significant developments are expected in technology based only on a boiler + steam
turbine. The contribution of stand-alone gas turbines is negligible. The competitiveness and
environmental performance of both CCGT and GT + HRSG technologies are definitely improving with
higher electricity efficiencies. The fact that there is less residual heat for heat applications if electrical
efficiency improvement is taking place in the future, does not present a problem as electricity has a
higher exergetic value and many installations with a higher electricity efficiency have a residual heat
surplus, even in CHP mode. Also, the domestic demand for electricity is expected to continue to grow
(see Chapter 1) while the demand for process and district heating will show a further decline as a
result of innovation-driven reductions in the energy needs per product in industry combined with an
increasingly greater share for electricity on the one hand and further insulation, additional energysaving measures (to e.g. lower EPC) and a decline in the number of degree days (ECN, 2005) on the
other hand. It should be noted that this excludes the potential which is currently being generated by
conventional means and can be supplied via process and district heating.
Fuel and Emissions
Many different gases can be used as fuel including natural gas, coal gas, blast furnace gas, syngas,
flare gas, and biogas. Liquid (clean) fuel can also be used with or without gas fuel (dual fuel). The socalled heavy fuels such as coal, heavy oil etc. must first be converted by the IGCC gasification
process into clean gas fuels before they can be used in the gas engine. This IGCC process is the
same as the one described in the previous section. A specially designed combustor is needed for fuel
such as coal gas, blast furnace gas and syngas because of their low LHV and hence the higher fuel
input requirements. A large volume of experience has been built up in this domain (Shell Per+,
Schwarze Pumpe, gas turbine GE MS6001B).
Though the technology for lowering emissions in smaller gas turbines is trailing slightly behind the
technology for the larger gas turbines, this is not a disadvantage because the emissions are
comparable in both cases.

153

Part 3

Other Developments
It is unclear just how far the technological and economic developments for smaller gas turbines (< 100
MWe) will run parallel with the developments for larger gas turbines. At both GE and Siemens the
electricity efficiency for industrial gas turbines is lower than for the large gas turbines. For example,
Siemens largest industrial CHP configuration (SGT-800, SCGT 45 MW, CCGT 64 MW) has a gas
turbine efficiency of 37% at a maximum total electricity efficiency of 54% while in the single shaft
(SSC5-4000F, SCGT 268 MW, CCGT 416 MW) the gas turbine efficiency is 39% at a maximum total
electricity efficiency of 58%: a difference of 2% SCGT and 4% CCGT respectively. The difference is
even greater in the case of smaller industrial CHP applications.
The same goes for GE industrial CHP applications. The efficiency of the larger industrial CHP Frame 6
configuration is between 32% (MS6001B, SC 42 MW, CC 64 MW) and 37% (MS6001C, SC 45 MW,
CC 67 MW) for a total electricity efficiency of between 49% and 54%, while the efficiency of the larger
F and H technology gas engines is between 37% (MS9001FA, SCGT 256 MW, CCGT 520 MWe) and
an estimated (no SCGTreport because of integrated steam-based blade-cooling) 39 - 40% (MS9001H,
CCGT 520 MWe) at a total efficiency of 57 - 60%, differences of 0 - 5% SCGT and 3 - 6 - 11% CCGT
respectively. It should be stressed that the lower total efficiency cannot be entirely attributed to a lower
gas turbine efficiency. Also, the configuration of the HRSG and the steam turbine is usually a bit
simpler in case of a power plant based on a smaller gas turbine. For example, single-pressure or
double-pressure HRSG are often used in industrial CHP instead of triple-pressure HRSG with reheat
in large combined cycles. The same applies to the configuration of the steam turbine. All of this is
done with a view to suppressing the CHP investment costs. The specific investment costs are higher
for CHP units than for large-scale units that run on gas turbines, thereby undermining the
competitiveness of CHP (Van Eck, 2007). If the developments in small-scale gas turbine technology
trail farther behind the developments in large-scale gas engine technology, then smaller-scale local
gas-turbine CHP solutions will become even less competitive in terms of costs and energy than they
are at present.
The higher levels of efficiency generated by gas turbines is also making it more attractive to build
simple and cheaper CHP plants consisting of a gas turbine and a HRSG (not a steam turbine) for
producing steam or hot water. This is particularly interesting if aero-derivative gas turbines are used
with very high simple cycle efficiencies of around 41% (GE, LM6000 PD Sprint) or 46% (GE,
LMS100PB, 98 MW) and lower exhaust gas temperatures of 446 and 417C respectively. Heavy duty
gas turbines with high efficiencies of 35 - 36% (GE, MS6001FA) are also used for this option. In
practice these plants are only deployed with full heat utilization and auxiliary boilers to absorb
fluctuations in demand (e.g. CHP Air Products, CHP Lyondel). The absence of a steam turbine
significantly lowers the specific investment costs.
Operation & Maintenance
Every year since 2004, ECN has calculated the non-profitable top for CHP projects in its MEP
recommendations (2004, 2005). These calculations also include the O&M costs for CHP plants, based
on consultations with market players and owners of CHP plants.
In its MEP recommendations for 2005 and 2006, ECN mentioned an O&M cost of 8.7/MWh in 2005
for a CCGT with a power output range of 50 - 70 MWe. For 2006 a differentiation was made for a large
and small CCGT. For a 250 MWe CCGT O&M cost of 4.7/MWh are mentioned and for a small, 80
MWe CCGT, 7.9/MWh.
For the model calculations the O&M cost of CCGT units < 250 MWe (for O&M cost for a CCGT > 250
MWe see Section 6.2.4) are set on 8.0/MWh for 2012 and 2025 .
The O&M mentioned for a GT + HRSG are 6.2/MWh for 2005 and 6.4/MWh for 2006. For the
model calculations the O&M cost are set on 6.5/MWh for 2012 and 2025.
Investment Costs
In the same recommendations for 2005 the investment cost mentioned for a 47 MWe CCGT with a
heat/power (H/P) ratio of 0.4 was 886/kWe, for a 81 MWe CCGT (H/P 0.8) it was 799/kWe and for
75 MWe CCGT (H/P 0.4) it was 683/kWe. In the ECN calculations for 2006 the mentioned specific
investment for a small CCGT was 949/kWe. It is clear that these investments are higher than for a
larger CCGT without heat.
For the model calculations the investment cost for a CCGT are set at 850/kWe in 2012 decreasing to
765/kWe (10% reduction) in 2025.

154

Chapter 6 Technological Trends in Supply Options

The recommendations for 2005 and 2006 set the investment costs of a GT + HRSG of 25 MWe at
979/kWe;
For the model calculations the investment cost for a GT + HRSG are set at 975/kWe in 2012
decreasing to 875/kWe (10% reduction) in 2025.
6.3.2

Small-Scale Combined Heat and Power (Gas Engines)

Technology development of gas engine leads to the raising the electricity efficiency. At present
efficiency of around 39% is attainable at 500 mg per Nm3 of NOx for gas engines of 330 kWe, 41.5%
for gas engines of around 1 MWe, and approximately 42.5% for gas engines in the range of 1.6-3.0
MWe (GE Jenbacher, 2004). Further rises are expected in these figures. Larger gas engines in the
range of 4 - 17 MWe attain even higher efficiencies of 46 - 47% at 500 mg per Nm3 of NOx (Wartsila,
2006). These turbines have been omitted from this study because they are used only to a limited
extent in the Netherlands. The competitiveness and the environmental performance of these
technologies are constantly improving with a higher electricity efficiency, but experience gained from
gas engines with heat distribution projects has shown that the operational efficiency for controlled
operation, on-off operation and operation without (full) utilization of residual heat is often lower in
practice than on paper.
Fuels and Emissions
A wide range of gases can be used as fuel, including natural gas, flare gas, biogas, landfill gas,
sewage gas, and special gases like mine gas and oxygas.
Gas engines generate relatively high NOx emissions for each unit of consumed fuel. Agreement was
reached in the Gothenburg Protocol that, at the end of 2007, a tighter limit of 250 mg per Nm3 with 5%
O2 (approx. 80g NOx per GJ) would be introduced for new gas engines. This is the limit that seems
technically feasible at the moment with engine tuning. These tighter environmental regulations will
lower the electricity efficiency by 1 - 1.5%. They will also lower the thermal efficiency and increase CO
and uncombusted hydrogen. The effects of defining the electricity efficiency still further are unclear.
Another option is a less tight adjustment with fine-tuning and flue-gas cleaning by selective catalytic
reduction (SCR), a technique which is used on many gas engines in the greenhouse sector, but which
requires extra investment and pushes up the running costs.
Other Developments
The horticultural sector aspires to a more environmentally friendly way of energy conversion in which
renewable energy play a major role. Two trends can be noticed (Energie-WEB, 2004):
In stead of producing the required temperatures for heating purpose by the degradation of
fossil fuel energy (and thereby lowering the exergy), transferring low energy heat, for example
energy of the sun, with use of heat pumps, towards required temperature levels;
Small-scale and large-scale greenhouse clustering (lit and unlit) with a large circle of
consumers, any residual heat could be shared and the operational times of the CHP could be
increased, thus saving energy and cutting costs (Cogen Projects, 2006).
The concept of the closed greenhouse is an example of the first trend. In this concept the air windows
of the greenhouse are permanent closed and the surplus of sun heat in the summer is stored in (warm
water) aquifers. This stored heat can be re-used for heating the greenhouse during the nights or
during the winter with use of heat pumps without further use of direct fired heat boilers. The cooled
down water from the heat pump is stored in (cold water) aquifers. This cold water is used for cooling
and for creating the right humidity atmosphere during summer conditions. The combination of the use
of cold water in the summer and warm water in the winter has to create a fixed temperature level in
the greenhouse, which does not require extra sources of heat production anymore.
Another innovative development is the so called energy producing greenhouse for which a pilot
project has been started in Bergerden recently (Roza, 2006). In this greenhouse the combination of a
strong isolating glass deck but with an efficient let through of ambient (sun) light, the very efficient
water/air heat exchanger called Fine Wire Heat Exchanger (Fiwihex), the permanent closing of the
windows (closed greenhouse) and the use of aquifers is combined. The working principle is as
described above, but without the use of a heat pump.

155

Part 3

Another technique that may be applied in the future is the Organic Rankine Cycle (ORC), which cools
a (waste or residual) heat flow of a relatively high temperature so that electricity can be generated. An
ORC that is linked to a gas engine can improve electricity efficiency by 5 - 8% (Cogen Projects, 2006)
potentially a good option for lit greenhouses as it makes for less surplus heat. Pilots are underway at
present but it is unclear when this technology will become available for gas engines on a large scale.
Operation & Maintenance
In its MEP recommendations for 2005 and 2006 ECN mentioned an O&M cost of 7.0/MWh for 2005
and 7.2/MWh for 2006 for a 1 MWe gas engine.
For the model calculations the O&M cost are set on 7.0/MWh for 2012 and 2025.
Investment Costs
In the same recommendations for 2005 and 2006 the ECN (2005) quoted a specific investment of
872/kWe for a 1 MWe gas engine.
For the model calculations the investment cost for a gas engine are set at 875/kWe in 2012
decreasing to 790/kWe (10% reduction) in 2025.
6.3.3

Small-Scale Combined Heat and Power ( WKK)

Micro-CHP, in a range of 1 - 5 kWe, is an extension of the idea of cogeneration to save primary


energy compared with separated generation of electricity and heat but then applied in single/multifamily homes or small office buildings. Micro-CHP systems, which operate in homes or small
commercial buildings, are normally driven by heat demand, delivering electricity as a byproduct and
will often generate more electricity than is instantly being demanded. This is also the case for the heat
production. In order not to destroy excess heat production, buffering of heat is required. A possible
pattern of electricity and heat production combined with buffering is shown in Figure 6.23.

Figure 6.23: Possible electricity and heat production of a micro-CHP unit (Source: EA Technology
Consulting, 2003)
Because of this operating model and because of the fluctuating demand for electricity in the system
buffering is a necessary option to let the micro-CHP produce electricity during the peak demand period
in the late afternoon while (partly) avoiding the use of the unit for peak demand only. Micro-CHP
systems can therefore achieve much of their savings (and thus attractiveness to consumers) through a
generate-and-resell or net-metering model where home-generated power exceeding the

156

Chapter 6 Technological Trends in Supply Options

instantaneous in-home needs is sold back to the electrical utility at the right moments. Another
advance is the local use of the generated electricity and so avoiding the common losses in the
transmission network from the source to the consumer. Another benefit of net-metering is the fact that
it is fairly easy to set up. The user's electrical meter is normally simply able to record electrical power
exiting as well as entering the home or business. As such, it records the net amount of power entering
the home.
While net-metering can be a very efficient mechanism for using excess energy generated by a microCHP system, it is not without its detractors. The first to consider is that while the main generating
source on the electrical grid is a large commercial generator, net-metering generators spill power to
the grid in a haphazard and unpredictable fashion when not optimized for electricity and heat
production during the right moments. Also the differences in electricity supply in summer and winter
are extreme so that in summer micro-CHP can hardly produce on times that peak electricity is
required or produces with a bad efficiency whilst the heat is not required. For a grid with relatively a lot
of micro-CHP users, probably design changes need to be made by the grid owner. From the
standpoint of the grid operator, these points present operational and technical as well as
administrative burdens. The first glance of micro-CHP only represents the consumers cost-savings of
not purchasing utility power versus the true cost of generation and operating the total system including
the grid.
Micro-CHP systems are currently based on several different technologies:

Reciprocating or piston engine

Stirling engines

Steam engines

Rankine cycle engines

Fuel cells.
This study focuses on micro-CHP based on Stirling engines used for single/multi-family homes
because this technology is outgrowing the demonstration.
Developments
Gasunie Engineering & Technology has been researching the potential of micro-CHP for home
installations since 1995. This research was prompted by, amongst others, the realization that the
energy potential of the High-Efficiency boiler is almost exhausted. Eventually, it can be replaced by
micro-CHP or heat pumps, hybrid boilers (HE boilers with a small electric heat pump), or solar combiboilers. The first trials with micro-CHP were disappointing, but since then models have appeared
which have inspired sufficient confidence. One such model is the Whispergen mk4 from the New
Zealand firm of Whisper Tech. It has electrical power of around 1 kW and thermal power of around 7
kW, based on a four-piston Stirling engine. In the spring of 2005 a field test was performed with
around 50 of these micro-CHPs. The results were encouraging. Since then, a successor has been
launched, the Whispergen m5, which has thermal power of 12 kWth and an extra combustor which, in
combination with an indirectly fired boiler, could replace the domestic combi-boiler. The English firm
Microgen has also developed a micro-CHP concept, a one-piston Stirling engine with electrical power
of around 1 kW and hot running water as an extra function. A field test was performed on this model in
the north of the Netherlands. A field test involving 1,000 units is expected in 2007.
The purpose of a virtual power plant is to remote-control and efficiently deploy multiple local electricity
generators. Gasunie wants to prove to local grid operators (who are often wary of conventional microCHP) in a demonstration project that centrally controlled (intelligent) micro-CHPs are not a threat, but
an opportunity, and that virtual power plants show that micro-CHP can play an active role in
maintaining the stability of local grids. The concept of the virtual power plant might also involve the
public in the energy issue; it could transform households from passive end users to active energy
producers (SenterNovem, 2006). The micro-CHP will therefore have to be started on the basis of the
demand for electricity. The idea is to store the generated heat as efficiently as possible so that it can
be used at a later date.

157

Part 3

Net Electrical Efficiency


The electricity efficiency of micro-CHP is in the range of 10 - 25% (see Figure 6.24). An avarage
electrical year efficiency of 15% is being attained by a study of MNP (2006) for the period until 2015.
Further improvements could raise this to 25% for 2015 2020 and 25 - 35% for 2020 - 2030, the
longer term (Ecofys et al, 2006).

Figure 6.24: Schematic of micro-CHP (Source: EA Technology Consulting, 2003)


In a presentation entitled The Future of CHP (2004) the NMP (National Institute for the Environment
and Public Health) assumes 20% electricity efficiency and 70% thermal efficiency for micro-CHP in the
range of 5 - 100 kW in 2040. The mentioned electrical efficiencies are theoretical percentages; the
actual (annual) figures are expected to be lower. For various reasons; for example, it is not always
possible to use all the generated heat (heat storage needed). Secondly, the demand for heat is very
low in the summer, so the plant does not need time to recover (the plant needs time to start up in
order to reach its operational temperature). The author of Micro CHP Doubt (Natta Journal, 2006) cites
efficiencies of between 5 and 15%.
In this study an electrical efficiency of 15% is applied for 2012 increasing to 25% in 2025.
Maintenance Costs
Microgen claims that the maintenance requirements for their micro-CHPs are the same as for
conventional boilers. As micro-CHP projects are still in the demonstration phase no further information
is available.
Investment Costs
Cogen Belgium has quoted investment costs of 1,300 42,000/kWe depending on the manufacturer
and whether a prototype is involved or machines which are going into production. ECN announced in
a study by the University of Utrecht (2001) for Enetac (co-operation ATAG, Eneco and ECN) that the
investment for a micro-CHP was approximately NLG 3,000 (approx. 1,350) more than for a central
heating boiler. An ECN study (2002) stressed that the difference between the price of a micro-CHP
and a central heating boiler must not exceed this amount. The same study says that the investment
costs for a High Efficiency boiler are between 1,800 and 2,500 but maintains 2,000 as a
guideline. This means that the investment costs for a micro-CHP should lie in the range of 3,350.
The sum applied in this study is 3,500 for a micro-CHP of 1 kWe.

158

Chapter 6 Technological Trends in Supply Options

6.4

Sustainable Options

6.4.1

Wind Energy

Wind turbine technology is moving fast. In the 1980s, wind turbines of 75 kW were the norm.
Nowadays wind turbines of 3,000 kW are no exception and turbines of 5,000 kWe are expected in two
or three years. The long-term aim is 10 - 15 MWe (Van Kuijk, 2005).
The price of wind turbines is dictated by the price of steel and carbon fiber. The manufacturers of wind
turbine blades are the main buyers of carbon fiber, which is needed in ever-greater quantities as
capacity is increased by enlarging the blades. The basic design for wind turbines is unlikely to change
much but the manufacturing techniques will change through developments in smart blades (for more
information see TU Delft and IEA RD&D), deformable surfaces, thermoplastics, and direct-drives (no
gear box). The hoisting and installation equipment also needs constant adjustment, not for the weight
but for the increase in the height of the rotor, needed for larger capacities (see Figure 6.25) especially
offshore.

Figure 6.25: Growth of WTG (Source: Chen, 2005)


Advances are also taking place in power electronics to meet the criteria for grid stability, controllability
of reactive power, and harmonics. If grid stability problems arise, the output of the wind turbine can be
more or less seamlessly regulated by pitch-angle control. The controllability of wind energy is getting
more important because of its ever-increasing contribution to the energy supply (Van Kuijk, 2005).
Operation & Maintenance cost for on-shore wind turbines will vary between 18 and 28/kW (1.6% of
the investment of 1150/kW according to ECN (2003). The European Wind Energy Information
Network mentioned a value of 31/kW for the Netherlands. Another ECN study (2004) mentioned a
range between 30 to 50 /kW. For this study 40/kW in 2003 is taken and with 1,800 equivalent full
load hours this leads to 22/MWh O&M cost. With an assumed fixed percentage of 3.5%
maintenance cost of the investment level, O&M cost decrease with the investment levels (see bellow)
to 37/kW ( 20/MWh) in 2012 and 32/kW ( 18/MWh) in 2025.
For offshore turbines the cost of O&M are significant higher, percentages of 4 - 4.5% of the investment
level are mentioned which leads to a range between 67 to 101/kW (ECN 2004).
For this study 90/kW in 2003 is taken and with 3,350 equivalent full load hours this leads to
27/MWh O&M cost and decreasing, because of lower future investment levels, to 79/kW in 2012 (
24/MWh) and 68/kW in 2025 ( 20/MWh).

159

Part 3

The investment costs for on-shore wind turbines are subject to wide variation. According to ECN
(ECN, 2003), they range between 895/kW and 1,150/kW and are higher for larger turbines with a
high mast and a large rotor, which are used in low wind climates. In this study for 2012 a decrease in
investment cost of approximately 10% is assumed compared with 1,150/kW and this leads to an
investment level of 1050 /kW and this is decreasing further with approximately 15% (ECN, 2004) to
900 /kW in 2025.
The offshore investment costs are much higher not only because the sea conditions call for firm
anchorage but also, of course, because of the connection with the mainland. Shell Venster (2005) has
stated that the investment costs for the 108 MW wind farm in the North Sea are over 200 million,
which works out at approximately 1,850/kW. An appendix by an ECN study (2006) cites investment
costs of 2,100/kW but predicts a drop to around 1,750/kW once experience has been gained.
In this study an investment of 1,750/kW is assumed for 2012 (decrease of 12.5% compared with
2,000 /kW in 2003) and a further decrease with 15% towards 1,500/kW in 2025.
6.4.2

Hydropower

The potential of traditional hydropower with reservoirs has more or less been realized in Europe. Only
in Iceland and a few other remote regions is there still untapped potential. Investigations have been
conducted into laying submarine pipelines from Iceland via England in order to transport the
unused hydropower and earth heat to the Netherlands where it can be used to generate electricity.
The technology for submarine pipelines has been around for years but the costs are prohibitive
(Kreuger, 2003). According to Kreuger, this technology becomes more feasible when compared with
offshore wind farms.
Hydropower with reservoirs or other storage basins would enable hydropower to be deployed in the
overall electricity system in a way that would level out the demand for plants fired by fossil fuel. This
would reduce the need for new building, provide more openings for CHP and wind energy, and lead to
more efficient business operations. Typical examples are the connections currently being installed
between Norway and the Netherlands, not to mention Lievenses alternative plans
(Begeleidingscommissie Voorstudie Plan Lievense, 1981) to fill the Markerwaard with water when
demand was low and the electricity supply was high. The excess capacity in electricity production
could then be used to raise the water level in the lake. When supplies were low and demand was high
the turbines would deliver more electricity. This plan was thought out as a means of coping with wind
energy problems. The electricity from wind turbines fluctuates with the wind speeds and has little to do
with demand. Lievense hoped to solve this problem by building reservoirs surrounded by a high dyke
upon which 400 wind turbines with a capacity of 1 - 1.5 MW would be built. The plan was never
implemented because aside from the landscape and ecological implications the costs were too
high and the safety risks too serious (a break in the dyke would flood Amsterdam).
To assess the effects of accumulation on the power supply system, calculations are performed for one
scenarios with use of the simulation model. These are based on a 1,500 MW accumulation through
pumps which replace Lievenses wind turbines (see also Section 6.4.5).
Hydropower also offers huge potential in the form of currents, waves, tides and differences in
temperature but this lies outside the scope of this study.
6.4.3

Solar Energy

Solar energy is, in principle, inexhaustible and is the greatest but least used source of energy on the
planet. Without sun, life would be impossible. If solar energy could be converted into all the required
forms of energy at the right time and place, the global energy problem would be solved. The following
applications are possible at present:
The production of hot water with solar boilers. This is now proven technology though there is
room for further improvement. This option is particularly suitable for sunny regions where the

160

Chapter 6 Technological Trends in Supply Options

sun is not too low on the horizon. It can also be used in moderate climates with less sun, but
then usually in combination with a back-up system to ensure that there is always enough heat.
Photovoltaic solar energy (PV) for generating electricity. The current technology is still
delivering low efficiency at high costs. The ROI time is long. Autonomous PV systems are
used in remote rural regions where there is no grid (streetlights, control boxes etc) or in
agriculture (drinking troughs for livestock). In such cases PV is cost-effective because it is a
substitute for a connection to the grid. Autonomous systems are also used at sea (illuminated
buoys). Often, they are combined with small storage systems for generated electricity.
Although PV applications for systems linked to the grid are still very expensive at present the
investment costs should eventually fall (Kema, 2006). If a way is found to store electricity or to
produce H2 with solar energy, then this option will have a deep impact on the energy situation.
There are concrete plans to generate solar energy by CSP (Concentrating Solar Power) in the
warmer (desert) regions of Africa where the sun shines all the time. A solar plant of 50 MW
has been built in Spain (van Nispen, 2006) and projects have been realized in the USA, albeit
at a high cost. A CSP solar plant consists of panels that concentrate the suns rays on one
spot. This can cause temperatures of 400 - 500C which can generate steam to drive a
turbine. Transporting electricity from this source to Europe is another issue. Whether it is
feasible depends on, amongst others, the transport losses and costs and the commodity prices
for electricity. ECN does not expect energy from the African sun before 2020.

According to KEMA (2006), the costs of PV systems for standard PV modules is 3/W. This is
expected to fall to 2/W in 2010, 1/W in 2020 and 0.5/W in 2030. A breakthrough in the
application of PV systems is expected after 2020 (KEMA, 2006).
6.4.4

Bio-Energy

Biomass Incineration / frementation and others


Bio-energy exists in many forms including wood, grass, straw, mulch, sludge, residues, compost, solid
recovered fuel, bio-energy bearers (as intermediate product) etc. Bio-energy can be converted into
electricity and/or heat by co-firing in existing plants or by thermal or bio-chemical conversion in standalone plants (see Section 4.6.4).
Waste Incineration
A new design of a waste incinerator is the so called High Efficiency waste incinerator build by the Afval
Energie Bedrijf (AEB) in Amsterdam in operation since 2007. The base principle of this waste
incinerator is the same as for conventional incinerators. The electrical efficiency of the High Efficiency
waste incinerator however is 30%, for the conventional incinerators this is approximately 22%. A
combination of high steam pressure, increase of the steam temperature and mainly the reheat of
expended medium pressure steam with saturated steam leads to this higher efficiency compared with
the conventional incinerators.
The boiler of the High Efficiency waste incinerator is designed for steam parameters of 440C/125 bar
but has the space for an extra bundle for superheating the steam towards 480C leading to a further
increase of the efficiency (AEB, 2006). ECN (2003) expects a limited technology development but
sees a trend towards large scale locations, more incineration lines per location or in increasing of
capacity of a single incinerator line. Producers can also choose for lower cost incineration in stead of
building a high efficiency waste incinerator (VROM, 2004).
For the model calculations however an efficiency of 30% is used for 2012 and 35% for 2025.
Operation & Maintenance cost for a waste incinerator will vary between 17 and 22/MWh according
to ECN (2003) but this is including the operation and maintenance for the waste incinerator part of the
installation. For this study it is assumed that operation and maintenance cost for the electrical part of
the installation are the same as for PCC installations i.e a total of fixed and variable cost of 7.0/MWh
used for 2012 and 2025.
The investment costs for waste incinerators vary depending on applied technology and size of scale.
According to CE (2001), the previous build AEB installation of 860 kTon/a waste treatment and 80

161

Part 3

MWe output, had an investment level of 428 M of which 23.9% was for the electrical part of the
installation. The Twence installation of 290 kTon/a and 30 MWe output, had an investment level of 235
M of which 23.8% was for the electrical part of the installation (both are price levels of 1990). More
recently the AEB high efficiency waste incinerator of 530 kTon/a and 54 MWe output was build for 450
M (Gemeente Amsterdam, 2006) and the fourth incinerator line of HVC Alkmaar of 150 kTon/a and
16.5 MWe was build for 110 M (HVC, 2005). With a power/waste ratio of 110 kWe/kTon and 24% of
the total invesetment for the electrical part of the installation, this leads to an investment for the
electrical part of the installation of approximately 1550/kWe assumed constant for 2012 and 2025.
6.4.5

Electricity storage

Electricity storage systems can decouple the timing of generation and consumption of electricity.
Specially for the integration of large amounts of wind power this could be a very useful system but also
for purpose of peak shaving these systems can be applied. Storage systems can be categorized in
small-scale systems and large scale systems (Hendriks, 2004). Storage technologies used for small
scale are super-conducting magnetic energy storage (SMES), super-capacitors and ultra-capacitors,
flywheel energy storage units and small scale batteries. These systems have power capacities of at
maximum several MWs and energy capacities up to 1 MWh. Storage technologies used for large
scale are pumped storage hydropower, underground pumped storage hydropower, compresses air
energy storage (CAES) and large-scale (flow) batteries. Energy capacity of these systems can reach
up to 50,000 MWh with power capacities up to 2000 MW for pumped hydro storage. In this study focus
is on large scale storage systems for the integration of large scale wind power
Pumped storage hydropower (PAC)
The plan Lievense (see also Section 6.4.2) is an example of a pumped storage hydropower system.
The latest variant of this system is a artificial energy island 15 km out of the Dutch cost. The idea is to
create, on a lower level, a lake in the North Sea which is capable to compensate the energy
production fluctuations of the several wind parks. This is done with the use of pump generators, to
much produced wind energy is used to pump water from the lake (lower storage basin) into the North
sea (upper storage basin). In case of shortage of electricity the water form the higher level sea is
directed via the pump generators back to the lake. The pump generators in that case produce
electricity. A feasibility study for a 2,000 MW artificial energy island, with an assumed energy rating of
50 GWh, is performed nowadays and first estimates for the investment cost amounts values between
1.5 and 2.0 billion excluding artificial island (Het Financieel dagblad, 2006). Specific investment
would in that case amount max 1,000 /kWe for the energy part, assumed is a total investment,
including artificial island of 1,500 /kWe.
Underground pumped storage hydropower (OPAC)
The working principle is the same as for a pumped storage hydropower plant but the main difference
is the difference in height of the reservoirs. Because in the Netherlands there are no large difference in
altitudes of water levels, the construction of a surface - and underground water reservoirs on a depth
of approximately 1.2 km has to be applied. The pump generators are placed close near the
underground reservoir. The amount of possible water content and differences in height between the
surface and underground reservoir is leading for the energy rating. The expected efficiency (regained
electrical energy) of a OPAC system is 75% (NEOM, 1986).
Compresses air energy storage (CAES)
Air is compressed and stored under pressure in a storage facility for compressed air for example, man
made rock cavern, salt cavern or porous rock, either created by water-bearing aquifers or a natural
depleted oil or gas reservoir (Hendriks, 2004). The pressurized air is used as combustion air to drive a
gas turbine generator installation. A compressor is not needed cause the air is pressurized already
and this leads to a quick starting installation. The largest CAES plant, 290 MW, is build in Germany
(Huntorf), typical power and energy rating are 100 - 300 MW and 1 - 10 GWh. Suitable sites in the
Netherlands could be found in empty gas fields or salt caverns but investment cost are expected to be
high.

162

Chapter 6 Technological Trends in Supply Options

Large-scale (flow) batteries


According to Hendriks (2004), Poly-sulphide bromide (PSB) technology has the largest energy storage
and power capacity compared to other flow batteries. It is expected that PSB systems will vary from 5
MW until 500 MW with typical energy ratings of 50 5,000 MWh and efficiencies of the system of
currently about 65% but technologies improvement will result in efficiencies up to 75%. Investment
cost for to existing testing projects of the PSB system were planned to be 1,400/kW but according to
KEMA (Hendriks, 2004) it is expected that investment cost will reduced to 800/kW in 2020.
6.5

Overview Technologies for Future Scenarios

6.5.1

Conventional Units

The development of Conventional and Combined Cycle technologies and the technical and economic
characteristics in the future are described in the previous section 6.2. The summarized characteristics,
used for scenario input in Chapter 9, are given in Tables 6.7 and 6.8 for respectively 2012 and 2025.
Conventional and Combined Cycle units in 2012 Characteristics used in model calculations
Type
Power
E-Efficiency (design) O&M
EPC+Other
PCC
1,200 MWe
45%
7.0/MWh
1,200/kWe
ABWR
1,465 MWe
36%
6.8/MWh
1,600/kWe
EPR
1,600 MWe
36%
6.8/MWh
1,875/kWe
PBMR
170 MWe
42%
5.0/MWh
1,400/kWe
IGCC
550, 900 MWe
47%, 46%
9.0/MWh
1,500/kWe
CCGT
480, 832 MWe
60%, 58%
3.5/MWh
700/kWe
Table 6.7: Technologies, characteristics and anticipated cost units first sight period (2012)
Conventional and Combined Cycle units in 2025 Characteristics used in model calculations
Type
Power
E-Efficiency (design) O&M
EPC+Other
PCC
1,200 MWe
53%
7.0/MWh
1,100/kWe
PCC with CCS
875 MWe
38%
13.0/MWh
1,900/kWe
ABWR
1,465 MWe
36%
6.8/MWh
1,450/kWe
EPR
1,600 MWe
36%
6.8/MWh
1,650/kWe
PBMR
300 MWe
45%
5.0/MWh
1,250/kWe
IGCC
600, 1,200 MWe 56%, 55%
7.0/MWh
1,200/kWe
IGCC with CCS
525, 1,050 MWe 44%, 43%
9.5/MWh
1,650/kWe
CCGT
500, 1,000 MWe 65%, 63%
3.5/MWh
600/kWe
CCGT with CCS
425, 850 MWe
55%, 53%
7.0/MWh
1,100/kWe
Table 6.8: Technologies, characteristics and anticipated cost units second sight period (2025)
The model implements the efficiency improvements between 2012 and 2025 in the period of
operation. In reality plants are being built over a long period of time which means that efficiencies are
sometimes higher for plants built after 2012 and lower for plants used in 2025. It is assumed that
efficiency differences are more or less neutralized.
6.5.2

Combined Heat and Power

The input for the scenarios in Chapter 9 use the developments of CHP technologies and
characteristics described in the Section 6.3, see Table 6.9 and 6.10 below (electrical efficiency with
heat delivery!).

163

Part 3

CHP units in 2012 Characteristics used in model calculations


Type
Power
E-Efficiency (average)
O&M
EPC+Other
CCGT
10 - 250 MWe
39%
8.0/MWh
850/kWe
GT + HRSG
10 - 100 MWe
33%
6.5/MWh
975/kWe
Gas Engines
0.3 - 3 MWe
36%
7.0/MWh
875/kWe
Micro CHP
1.0 kWe
15%
NA
NA
Table 6.9: Technologies, characteristics and anticipated cost units first sight period (2012)
CHP units in 2025 Characteristics used in model calculations
Type
Power
E-Efficiency (average)
O&M
EPC+Other
CCGT
10 - 250
43%
8.0/MWh
765/kWe
GT + HRSG
10 - 100
36%
6.5/MWh
875/kWe
Gas Engines
0.3 - 3 MWe
39%
7.0/MWh
790/kWe
Micro CHP
1.0 kWe
25%
NA
3,500/kWe
Table 6.10: Technologies, characteristics and anticipated cost units second sight period (2025)
6.5.3

Sustainable Options

The input for the scenarios in Chapter 9 use the development of sustainable technologies and
characteristics described in the section 6.4, see Table 6.11 and 6.12 below. As efficiencies are not
really applicable to this kind of application, the focus is on investment costs for wind, solar and waste
incinerator technologies. Developments in hydro (except for plans concerning water reservoirs) are not
relevant in the Netherlands and bio energy is mostly converted in existing or future conventional and
CHP units (see previous sections).
Sustainable options in 2012 Characteristics used in model calculations
Type
Power
O&M
EPC+Other
Wind onshore
1 - 5 MW
20/MWh
1,050/kWe
Wind offshore
1 - 5 MW
24/MWh
1,750/kWe
Hydro
NA
NA
NA
Solar
NA
NA
2/W
Bio energy
NA
NA
Inc., Fermt.,Others NA.
15 - 100 MW
7/MWh
1,550/kWe
Waste Incinerator
Table 6.11: Technologies and anticipated costs sustainable options first sight period (2012)
Sustainable options in 2025 Characteristics used in model calculations
Type
Power
O&M
EPC+Other
Wind onshore
1 - 10 MW
18/MWh
900/kWe
Wind offshore
1 - 10 MW
20/MWh
1,500/kWe
Hydro
NA
NA
NA
Solar
NA
NA
1/W
Bio energy
NA
NA
Inc., Fermt.,Others NA
15 - 100 MW
7/MWh
1,550/kWe
Waste Incinerator
Energy storage
100 - 2,000 MW 1% investment/a
1,500/kWe
Pumped Hydro
5 - 500 MW
1% investment/a
800/kWe
PSB flow battery
Table 6.12: Technologies and anticipated costs sustainable options first sight period (2025)

164

Chapter 6 Technological Trends in Supply Options

6.6

Summary

The main technological developments of the three main categories conventional and combined cycle
technology, combined heat and power technology and sustainable production technology (including
electricity storage) of the supply options are described in detail. These development are translated
in technical and economic characteristics for the future scenarios in 2012 and 2025 and are
summarized in Tables. The characteristics serve as direct input for the model of the Dutch electricity
supply.

165

References

References
Afval Energie Bedrijf AEB (2006), Mr waarde uit afval, HR Centrale, De nieuwe standard voor het
produceren van duurzame energie, metalen en bouw materialen uit stedelijk afval [Making More of
Waste. High Efficiency Power Plant: the new standard for producing sustainable energy, metals and
building materials from urban waste]
Areva (2005), EPR, Brochure
Areva (2005), Olkiluoto 3, A Turnkey EPR Project (European Pressurized Water Reactor), Presskit
Areva (2007), Finnish EPR Olkiluoto 3, The worlds first third-generation reactor now under
construction
Australian Uranium Association (2005), Advanced Nuclear Power Reactors, Nuclear Issues Briefing
Paper 16, December 2005
Basler, B. (2001), Power Plant Emission Reduction Potentials Achievements and Future Reduction
Potentials, Paper presented at the 18th WEC in Buenos Aires, Argentina, October 2001
Ber, J.M. (2000), Combustion Technology Developments in Power Generation in response to
Environmental Challenges, progress in energy and Combustion Science 26 (2000) 301- 327
Begeleidingscommissie Voorstudie Plan Lievense (1981), Windenergie en waterkracht: rapport van de
begeleidingscommissie Voorstudie Plan Lievense, staatsuitgeverij s Gravenhage, mei 1981, ISBN 9012-03541-4 [Wind Energy and Hydro-Power: report by the Steering Committee for the Lievense
preliminary plan]
Black, C.R. (2003), Future Options for Generation of Electricity from Coal, Prepared Witness
Testimony for the Committee of Energy and Commerce, June 24, 2003
Blankinship, S. (2006), Coal Gasification: Players, Projects, Prospects, Power Engineering July, 2006
Boer, B. (2004), Feasibility of a High Temperature Nuclear Reactor as Back-up System for Wind
Power, MSc Thesis, Report no. ET 2150, Delft University of Technology
Botha, F. (2004), Overview of the Fluidized Bed Combustion Process and Material, Illinois Clean Coal
Institute, paper presented at Technical Interactive Forum May 4-6, 2004 Harrisburg, Pennsylvania
Bugge, J. (2003), Advanced 700C PF Power Plant, Presentation Tech-wise A/S
Bugge, J., Kjr, S., Blum, R. (2003), High-Efficiency Coal-Fired Power Plants, Development and
Perspectives, Paper Elsam Denmark
CE (2001), Elektriciteit uit AVIs, Eindrapport [Electricity from Waste Incinerators]
CE (2005), Welke nieuwe elektriciteitscentrale in Nederland, Report No. 4800003493 [Which new
electricity plant in the Netherlands?]
Chen, Z. (2005), Issues of Connecting Wind Farms into Power Systems, Transmission and
Distribution Conference and Exhibition: Asia and Pacific, 2005 IEEE/PES Volume, Issues, 2005
Page(s) 1-6
Coca, M.T. et al. (1998), Coal Gasification. Conception, Implementation and Operation of the Elcogas
IGCC Power Plant, Paper presented at the 17th WEC in Houston, Texas, USA, September 1998
COGEN Europe (2003), Micro-CHP needs specific treatment in the European Directive on
Cogeneration

166

References

Commission for Energy Regulation CER (2002), Best New Entrant Price 2003, Consultation Paper
CER/02/151 4 October 2002
Damen, K.J. (2007), Reforming Fossil Fuel Use, The Merits, Costs and Risks of Carbon Dioxide
Capture and Storage, 2007, blz.238.
Damen, K.J., Van Troost, M., Faaij, A.P.C, Turkenburg, W.C. (2006), A comparison of Electricity and
hydrogen production systems with CO2 capture and storage, Part A: Review and selection of
promising conversion and capture technologies. Progress in Energy and Combustion Science, 2006;
32 (2)
Department of Trade and Industry DTI (2000), Technology Status Report, Flue Gas Desulphurization,
Cleaner Coal Technology Program
EA Technology Consulting (2003), Micro CHP in Europe, Presentation by J. Harrison
ECN (1995), Prospects for Energy Technologies in the Netherlands, Volume 2, Technology
Characterizations and Technology Results, ECN-C--95-039
ECN (2002), Micro-Warmtekrachtsystemen voor de energievoorziening van Nederlandse
huishoudens, ECN-C--02-006 [Micro-CHP Systems for Supplying Energy to Dutch Households]
ECN (2003), Kosten Duurzame Elektriciteit, Windenergie op land, ECN-C--03-074/A [Costs of
Sustainable Electricity: wind energy on land]
ECN (2003), Kosten Duurzame Elektriciteit, Afvalverbrandingsinstallaties, ECN-C--03-074/E [Costs of
Sustainable Electricity: waste incineration plants]
ECN (2003), Kosten Duurzame Elektriciteit, Kleinschalige waterkracht, ECN-C--03-074/G [Costs of
Sustainable Electricity: small-scale hydro-power]
ECN (2004), Onrendabele toppen van duurzame elektriciteitsopties, ECN-C--04-101 [Unprofitable
Tops of Sustainable Electricity Options]
ECN (2004), Advies WKK MEP-tarief 2004, ECN-C--04-049 [Recommendations CHP MEP Rate 2004]
ECN (2005), Advies WKK MEP-tarief 2005, ECN-C--04-111a [Recommendations CHP MEP Rate
2005]
ECN (2005), MEP-advies WKK 2006, ECN-C--05-102 [Recommendations CHP MEP Rate 2006]
ECN (2005), Referentieramingen energie en emissies 2005-2020, ECN-C--05-018 [Reference
Estimates for Energy and Emissions 2005-2020]
ECN (2006), Optiedocument energie en emissies 2010/2020, ECN-C-05-105, Appendix CO2-ENE-28,
Versie 13 maart 2006 [Document on Energy and Emissions Options 2010/2020, Version March 13,
2006]
Ecofys et al. (2006), Technisch energie- en CO2 besparingspotentieel van micro-WKK in Nederland
(2010-2030) [Technological energy and CO2 savings potential of micro-CHP in the Netherlands (20102030]
Ellul, C. (2005), Zero Emission Fossil Fuel Power Plant Technologies: a realistic future? Engineering:
Clean & Green, The XIV Annual Engineering Conference, Venue April 14, 2005
Energie-WEB (2004), Glastuinbouw in een duurzaam regionaal energienetwerk [Market Gardening in
a Sustainable Regional Energy Grid]
Energy Information Administration EIA (2006), International Energy Outlook 2006, Chapter 6
Electricity, 63-70

167

References

E.on UK PLC (2006), Performance of Whispergen Micro CHP in UK Homes


E.on Benelux (2006), E.ON kiest Maasvlakte voor bouw grote electriciteitscentrale, Publication date
08.03.2006 [E.ON chooses Maasvlakte as building site for large power plant]
Gasteiger, G. (2006), Perspectives for Economic and Climate-Friendly Power Generation, Forum II
Fossil Fired Power Generation Position of Technology Suppliers, Alstom Power AG Presentation on
World Energy Dialogue (WED), Hanover, April 26, 2006
GE Energy (2004), New High Efficiency Simple Cycle Gas Turbine GEs LMS100, GER-4222A
(06/04)
GE Energy (2006), Gas Turbine and Combined Cycle Products
GE Jenbacher (2004), Jenbacher Type 6, Gas Engines Product Brochure
GE Jenbacher (2005), Jenbacher Type 4, Gas Engines Product Brochure
GE Power Systems (2000), Combined - Cycle Product Line and Performance, GER-3574G (10/00)
GE Power Systems (2000), Power Systems for the 21st Century- H Gas Turbine Combined - Cycles,
GER-3935B (10/00)
GE Power Systems (2000), GE IGCC Technology and Experience with Advanced Gas Turbines, GER
4207 (10/00)
GE Power Systems (2001), Combined - Cycle Development Evolution and Future, GER-4206 (04/01)
GE Power Systems (2001), Advanced Technology Combined Cycles, GER-3936A (05/01)
GE Power Systems (2001), Creating Owners Competitive Advantage Through Contractual Services,
GER-4208 (05/01)
Gemeente Amsterdam (2006), Nieuwbouw Hoogredement centrale door Afvalenergiebedrijf
Amsterdam, Brief Aan de leden van de Raadscommissie KSB [New high-efficiency power plant
through Amsterdam Waste & Energy Company]
Geosits, R.F., Schmoe, L.A. (2005), IGCC The Challange of Integration, Proceedings of GT2005
ASME Turbo Expo, June 6-9, 2005, Reno-Tahoe, Nevada, USA
Geveke Motoren (2006), Caterpillar Gas Generator Sets, Gas Engines Product Brochure
Golay, M.W. ( ), Sustainable Energy, Generation IV and the Future of Nuclear Power
Greenpeace (2006), Fossil Fuel, Power-Plant Forum, Presentation on World Energy Dialogue,
Hanover, April 26, 2006
Gross, M., et al. (2000), Gasification of Residue as Source of Hydrogen for Refining Industry in India,
2000 Gasification Technolgies Conference, San Fransisco, California USA, October 8 - 11, 2000
Harrison, J., et al. (2001), Domestic CHP, What are the potential benefits?
Hellfritsch, S. (2007), et al., Modern Coal-Fired Oxyfuel Power Plants with CO2 Capture Energetic and
Economic Evaluation
Hendriks, R. (2004), Electricity Storage: a solution for wind power integration?, MSc Thesis, Report
no. ET 2150, Delft University of Technology
Het Financieel Dagblad (2006), Energieopslag: Kunstmatig eiland vangt schommelingen op, Special
Energie, 20 - 21, R. op het Veld [Energy Storage: artificial island offsets fluctuations]

168

References

Huisvuilcentrale HVC (2005), HVC opent vierde verbrandingslijn, press report [HVC opens fourth
incineration line]
IAEA (2006), Description of Natural Circulation and Passive Safety Systems in Water-Cooled Nuclear
Power Plants, CRP on Natural Circulation Phenomena, Modeling and Reliability of Passive Systems
that Utilize Natural Circulation, Draft
Interview with Professor G. van Kuijk (2005), TU Delft Lucht en Ruimtevaart, afdeling Windenergie
IPCC (2001), Climate Change 2001: The Scientific Basis, Cambridge, UK: Cambridge University
Press
IPCC (2005), IPCC Special Report on Carbon Dioxide Capture and Storage, prepared by Working
Group III of the Intergovernmental Panel on Climate Change, Cambridge University Press
Japan Nuclear Energy Safety Organization (2005), Current Status of Nuclear Facilities in Japan
Kadak, A.C. (1998), Economic Analysis of the Modular Pebble Bed Reactor, Massachusetts Institute
of Technology
Kadak, A.C. (2005), A Future for Nuclear Energy Pebble Bed Reactors, Int. J. Critical
Infrastructures, Vol. 1, No.4, pp 330-345
Kather, A. (2006), Worldwide R&D Activities for Climate-Friendly Fossil Power Generation in Fossil
Fuel Power Plants, Presentation on World Energy Dialogue, Hanover April 26, 2006
KEMA (2006), State of the art studie techniekontwikkeling voor de elektriciteitsvoorziening, 40560104TDC 06-55631A [State-of-the-Art Study on Technological Development for the Electricity Supply]
Kikstra, J.F. (2001), Modelling, Design and Control of a Cogeneration Nuclear Gas Turbine Plant, PhD
thesis, Delft University of Technology, ISNB: 90-9014631-9
Kitto, J.B. (1996), Developments in Pulverized Coal-Fired Boiler Technology, presentation at Missouri
Valley Electric Association Engineering Conference, Kansas City April 10 - 12, 1996
Korobitsyn, M.A. (1998), New and Advanced Energy Conversion Technologies: analysis of
cogeneration, combined and integrated cycles, PhD thesis, Twente University of Technology
Kreuger, F.H. (2003), Waar staan we met windenergie?, Nationaal Kritisch Platform Windenergie
[Where do we stand on wind energy?]
Ladwig, M. (2005), Introduction to Advanced Gas Turbine Technology, Alstom Power Presentation at
GCC-EU Seminar Natural Gas Advanced Technologies Realities and Prospects, Feb 7-8, 2005
Doha, Qatar
Lysen, E. (2007), CO2 afvang en opslag : kansen voor Nederland, Presentatie tijdens Congres :
Emissiehandel in een veranderend klimaat, Ede 15 May 2007 [CO2 Capturing and Storage: chances
for the Netherlands - Presentation]
Matzner, D. (2004), PBMR Project Status and the Way Ahead, 2nd International Topical Meeting on
High Temperature Reactor Technology, Beijing, China, September 22-24, 2004
Milieu en Natuurplanbureau MNP (2004), De toekomst van WKK en andere energieopties in de
energievoorziening: Schoon en Duurzaam?, Presentation held at Symposium Cogen 22 October
2004, Zeist, the Netherlands [The Future of CHP and other Energy Options in the Energy Supply
System: clean and sustainable?]
Milieu en Natuurplanbureau MNP (2006), Micro-warmtekracht en de virtuele centrale, Evaluatie van
transities op basis van systeemopties, MNP Report 500083003/2006 [Micro CHP and the Virtual
Power Plant; evaluation of transitions on the basis of system options]

169

References

Ministerie van VROM (2004), Voortgangsrapportage Landelijk afvalbeheerplan (LAP),


VROM/DGM/SAS BAOO May 2004 [Progress Report on the National Waste Management Plan]
Murase, A., et al. (2005), The Development of the Next Generation ABWR (AB1600), Proceedings of
Global 2005, Tsukuba, Japan, Oct 9-13, 2005 Paper No. 194
Nattas Journal, 2006, Micro CHP Doubts, Renew, Issue 161 May-June 2006
Nelson, R. ( ), Wind Energy, Resource Advantages and Constraints, Engineering Extension
Nelson, V. (2006), Wind Energy and Wind Turbines for Windy Land Owners, Alternative Energy
Institute, West Texas, A&M University
Nederlandse Energie Ontwikkelings Maatschappij B.V. NEOM (1986), Opslag van elektriciteit voor
morgen [Electricity Storage for tomorrow]
Novem (2006), NEO*, Nieuw Energie Onderzoek in Praktijk [New Energy Research in Practice]
Nuon (2006), Nuon Magnum, Multi-fuel Power Plant, Brochure
Oka, Y. (2001), Towards Nuclear Renaissance, A Perspective of Nuclear Energy and its Research,
Int. Conference E. Fermi and Nuclear Energy, Pisa, Italy, October 15-16, 2001
Parkinson, G. (2004), OEMs getting ready for coal gasification, Turbomachinery International
May/June 2004
Power (2003), Baglan Bay Power Plant, Cardiff, Wales, UK, Power July/August 2003
Ratafia-Brown, J, et al. (2002), An Environmental Assessment of IGCC Power Systems, Presented at
the 19th Annual Pittsburgh Coal Conference, September 23 - 27, 2002
Riahi, K, et al (2004), Technological Learning for Carbon Capture and Sequestration Technologies,
Energy Economics, 26(4) 539-564
Roza, C. (2006), Kas als energiebron, Keerpunt en katalysator, InnovatieNetwerk / Stichting Innovatie
Glastuinbouw [The Greenhouse as an Energy Source, Turning Point and Catalyst]
Santosh, K, et al. ( ), Simultaneous Removal of H2S and NH3 Using Metal Oxide Sorbents
Seabright, J., et al. (2001), Environmental Enterprise: carbon sequestration using Texaco gasification
process, presented at: First National Conference on Carbon Sequestration, May 14-17, 2001,
Washington D.C.
Siemens (2001), BENSON Boilers for Maximum Cost effectiveness in power plants
Shell Venster (2005), Wind aan Zee, Shell Venster, September/October 2005, p 6-9 [Wind at Sea]
Society of British Gas Industries (2004), Kick-Starting Micro CHP via the Energy Efficiency
Commitment
Toshiba Corporation, General Electric Company et al. (2005), New Nuclear Power Plant Licensing
Demonstration Project, ABWR Cost/Schedule/COL Project at TVAs Bellefonte Site
The University of Chicago (2004), The Economic Future of Nuclear Power, Study Report
Universiteit Utrecht (2001), Micro-WKK: meer dan koppeling van Kracht en Warmte alleen [Micro
CHP: more than the combining of power and heat]
Uranium Information Centre (2006), The Economics of Nuclear Power, Briefing Paper 8, April 2006

170

References

U.S Doe Nuclear Energy Research Advisory Committee, Generation IV International Forum (2002), A
Technology Roadmap for Generation IV Nuclear Energy Systems
Van Buijtenen, J.P. (1997), Gasturbine: motor voor hoog rendement en lage emissies,
Energietechniek 11, Volume 75, November 1997 [Gas Turbine: engine for high efficiency and low
emissions]
Van Eck, T. (2007), A New Balance for the Energy Sector: no longer a puppet in the hands of
technology, public interests and markets, ISBN 9789078889014
Van Nispen, P. (2006), Concentrating Solar Power, Interconnectiecapaciteit vormt struikelblok,
Energie Magazine, No. 3, p 16-18, May 2006
VGB (2004), CO2 Capture and Storage, a VGB report on the state of the art
Voges, K. (2004), Innovation and Efficiency, global energy report, official publication of the 19th WEC
Congress 2004, Sydney, Australia
White, A. (2005), Dawn of a New Era, WNA Symposium September 8, 2005
World Coal Institute (2003), Clean Coal - Building a Future through Technology
Zuideveld, P., et al. (2003), Overview of Shell Global Solutions Worldwide Gasification Developments,
Gasification Technologies 2003, San Francisco, California USA, October 12 - 15, 2003

Internet Addresses
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.

http://www.areva-np.com
http://www.australiancoal.com.au
http://www.coal21.com.au/
http://www.co2-cato.nl
http://www.eatechnology.com
http://www.eia.doe.gov/oiaf/ieo/
http://www.eia.doe.gov/cneaf/nuclear/page/analysis/nucenviss2.html
http://www.gepower.com
http://www.gtb.cogenprojects.nl
http://www.iea-coal.org/site/ieacoal/home
http://www.iter.org
http://www.nuc.berkeley.edu/designs/abwr/abwr.html
http://www.ol3.areva-np.com
http://www.pbmr.com
http://www.power.alstom.com/
http://www.powergeneration.siemens.com
http://www.power-technology.com/projects/sanlorenzo/
http://www.power-technology.com/projects/isab/specs.html
http://www.rwe.com/generator.aspx/konzern/fue/strom/co2-freieskraftwerk/language=en/id=268960/co2-page.html
http://www.senternovem.nl/eos/projecten
http://www.toshiba.co.jp/nuclearenergy/english/business/reactor/abwr.htm
http://www.wartsila.com/
http://ieawind.org
http://world-nuclear.org/info/

171

References

172

Part 4

Dutch Electricity Demand and Supply Scenarios and


Simulation Model

174

Chapter 7 Scenarios for the Dutch Electricity Supply

Chapter 7
7

Scenarios for the Dutch Electricity Supply


7.1

Introduction

The central role that scenarios for the Dutch electricity supply play in this study is described in detail in
the introduction in which the research objective is defined as follows:
Which scenarios can be realized in the Dutch energy supply with a view to the trends in supply
and demand of electricity, heat and fuel supplies? What effects do they have on the three
domains energy saving/reduction of CO2 emissions, economic efficiency, and the retention of a
high level of security of supply? And what measures from government and utilities are needed
to support them?
The scenarios developed in this chapter are used as an input for calculations made with the model
(see Chapter 8 and 9) which determines quantitatively the economics, the environmental effects, and
the balance between supply and demand. The outcome of the scenarios is then analyzed for a given
period in order to draw conclusions about the differences between the three above-mentioned
objectives. Combined with the analyses in the previous parts of this thesis, these conclusions enable
the appropriate measures from government and utilities to be developed for the scenarios.
This chapter starts by defining what a scenario is and then describes the method used to develop
scenarios. Before we started developing the scenarios for this study, we first researched the literature
to establish how other parties in the Netherlands view scenarios for the Dutch electricity/power supply
and how the scenarios are designed.
The scenarios used in this study are designed in stages. First, the driving forces are determined, and
then their uncertainties. The results of a number of the driving forces' uncertainties are used to create
a matrix of extremes. Combined with definitions of their horizons, some four to six scenarios can be
developed and formulated.
7.2

Scenarios

7.2.1

Definitions

What exactly is a scenario? Below are a few of the definitions found in the literature:
In its special report Emission Scenarios (2000) the IPCC defines scenarios as follows:
Scenarios are images of the future, or alternative futures. They are neither predictions nor forecasts.
Rather, each scenario is one alternative image of how the future might unfold. A set of scenarios
assists in the understanding of possible future developments of complex systems. Some systems,
those that are well understood and for which complete information is available, can be modeled with
some certainty, as is frequently the case in the physical sciences, and their future states predicted.
However, many physical and social systems are poorly understood, and information on the relevant
variables is so incomplete that they can be appreciated only through intuition and are best
communicated by images and stories.
In People and Connections, Global Scenarios to 2020 (2002), Shell defines scenarios as follows:
Scenarios are a tool for helping managers plan for the future or rather for different possible futures.
They help us focus on critical uncertainties: the things we dont know about that might transform our
business, and the things we do know about that might involve unexpected discontinuities. Scenarios
are alternative stories of how the world may develop. They are not predictions, but credible, relevant,

175

Part 4

and challenging alternative stories that help us explore 'what if' and 'how.' Their purpose is not to
pinpoint future events, but to consider the forces that may push the future along different paths.
The Dutch General Energy Council (AER, 2004) defines scenarios as follows: Scenarios are
internally consistent, somewhat extreme images of the future that are generally determined by
external factors, the goal of which is to enable us to evaluate our own policy from a number of different
possible developments. Scenarios are not predictions they display the extremes of a possible
future.
It is clear from each of the definitions that scenarios are not used to predict the future but to describe a
range of possible futures. A number of scenarios is used to gain an understanding of and insight into
extremes. Scenarios formulated with help of models specify the future in quantitative terms and, if fully
documented, they are also reproducible. Sometimes scenarios are less quantitative and more
descriptive and in a few cases they do not involve any formal analysis and are expressed in qualitative
terms (Figure 7.1). In this study, the scenarios selected to determine the future development of the
Dutch electricity supply will be calculated quantitatively with a model to determine the economics, the
environmental effects, and the balance between supply and demand. The quantitative analysis of the
different scenarios for selected periods enable conclusions to be drawn about the differences between
the three objectives described in this chapter's introduction.

Figure 7.1: Quantitative and qualitative analysis in combination with scenarios (Source: IPCC, 2000)
7.2.2

Scenario Development

Now that insight in scenario definitions has been given, scenario development (Wright, 2004) is the
next step, sometimes referred to as scenario writing (Weingand, 1995) or scenario building (McNair,
2000). In this study the term scenario development is used.
McNair (2000) specifies that the initial step of scenario building is determined by the drivers that will
propel future change. Drivers are the underlying factors in current trends, such as population growth or
decline, technological development and diffusion, or human factors like the will to power. Changes in
drivers result in changes in trends, which, in turn, result in changes in the human environment.
For Wilkinson (2004), too, the drivers are central: Since scenarios are a way of understanding the
dynamics shaping the future, we next attempt to identify the primary 'driving forces' at work in the
present. These fall roughly into four categories:
Social dynamics quantitative, demographic issues like migration, aging, and fertility, softer
issues of values, lifestyle, demand, or political energy;

176

Chapter 7 Scenarios for the Dutch Electricity Supply

Economic issues macroeconomic trends and forces (international trade flow and exchanges)
shaping the economy as a whole, microeconomics dynamics (competitors) and forces at work,
on or within the company itself, prosperity grow;
Political issues electoral (policy chosen government), legislative (tax policies), regulatory
(political instruments) and litigative (influences courts);
Technological issues direct (the speed new technology comes available), enabling (influence
new technologies) and indirect.

The uncertainty of some primary forces is also explained: Once these forces are enumerated, we can
see that from our own viewpoint, some forces can be called 'predetermined' not in a philosophical
sense, but in that they are completely outside our control and will play out in any story we tell about
the future. After we identify the predetermined elements from the list of driving forces, we should be
left with a number of uncertainties. We then sort these to make sure they are critical uncertainties. A
critical uncertainty is an uncertainty that is key to our focal issue. Our goals are twofold we want to
better understand all the uncertain forces and their relationships with each other. But at the same time,
we want the few that we believe are both most important to the focal issue and most impossible to
predict to float up to the surface.
Finally, Wright (2004), for whom driving forces are also key: It involves identifying the driving forces,
their degree of uncertainty and potential impact on the organization, the capture of the essence of the
scenario end state, and a fleshing out of storylines of the narratives. Wright also believes that it is
important to identify the areas with the highest level of uncertainty and impact on the issue or question
around which the scenarios are being constructed.
Once the driving forces have been determined, their degree of uncertainties must be determined.
Wilkinson says: At first, all uncertainties seem unique, but by stepping back we can reduce bundles of
uncertainties that have some commonality to a single spectrum, an axis of uncertainty. If we can
simplify our entire list of related uncertainties into two orthogonal axes, then we can define a matrix
(two axes crossing) that allows us to define four different, but plausible, quadrants of uncertainty, Each
of these far corners is, in essence, a logical future that we can explore." The matrix can be used to
create and work out narratives.
Now that we have dealt with the definitions, we can take a look at the schematic overview of scenario
development (see Figure 7.2):
Driving Forces Uncertainties Matrix Narratives Implications
Figure 7.2: Schematic overview of scenario development
7.2.3

Scenarios for the Dutch Energy and Fuel Supply in the Literature

Before we start developing scenarios for this thesis, we need to look at the literature to find out what is
known about scenarios for energy supply in the Netherlands. This literature survey, the analyses from
Part 2 and 3 and external influences on the energy supply system as described by Van Eck (2007) will
help to determine the driving forces, the uncertainties, and the matrix that will be used to develop a
number of scenarios. The most important literature is listed below:
General Energy Council, Carefully Upstream
In their advise for policy options for the Dutch electricity market, the General Energy Council (AER,
2004), bases its scenarios on the size of the market and the degree of effective competition (see
Figure 7.3).

177

Part 4

EFFECTIVE COMPETITION ON NW-EU LEVEL


SELF SUPPORTING
IN OWN COUNTRY

OLIGOPOLIE WITHOUT EFFECTIVE


COMPETITION ON NW-EU LEVEL

Effective
competition
Market share

Figure 7.3: Scenarios for the electricity market according to the General Energy Council (Source:
Algemene Energieraad, 2004)
The council believes that different scenarios in which there is more or less effective competition and
more or less supranational collaboration may unfold in the next decades:
Scenario I: EU+, with effective competition in a North-West European market. In this scenario, the
most practically feasible ideal is realized: the creation of one wholesale market in North-West Europe
with sufficient (more than five) players, liquidity, and transparency.
Scenario II: National self-sufficiency. In this scenario, bad experience with imports and black-outs
leads the member states to conclude that it is not desirable to be dependent on others for the supply
of electricity. Everyone wants to have sufficient generation capacity in their own country. Electricity is
imported or exported on very limited scale, but the concept of an international market has been
abandoned.
Scenario III: EU-, an oligopoly without effective competition in a North-West European market. In this
scenario, the geographic area is successfully expanded, but an effective competitive market structure
is not created. There are not enough players; there is market power and market domination.
Regulation and supervision remain fragmented, preconditions differ per country.
TenneT, Capacity plan 2003 - 2009
In its capacity plan (Tennet, 2002), TenneT uses socioeconomic developments and environmental /
sustainability aspects as design dimensions to select the following four long-term scenarios:
"Green Europe" scenario: A pan-European free market for electricity with a substantial share of
renewable energy emerges in an open society which is economically highly dynamic at a global level.
"Borderless Europe" scenario: Here too, long-term economic prosperity paves the way for a panEuropean free market for electricity in which renewable energy, however, attracts little interest.
"Self-sufficient region" scenario: Problems in the energy supply shatter the belief in deregulation and in
a free market. The problems also result in low economic dynamics, which weakens the support for the
further development of a sustainable society.
The "Regional sustainability" scenario resembles that of the self-sufficient region with the exception of
the attention paid to the environment and to sustainability. Renewable resources and a sustainable
environment are important social prerequisites.

178

Chapter 7 Scenarios for the Dutch Electricity Supply

Ministry of Economic Affairs, Energy and Society in 2050


In Energy and Society in 2050 (Ministry of Economic Affairs, 2000), four world views are outlined for
the Dutch long-term energy supply, which are assessed against quality criteria for security of supply,
economic efficiency, and penetration of renewable energy in 2050. The four world views are:
Free trade: Economics and money predominate without national barriers.
Isolation: Financial profit predominates within national/regional borders.
Big solidarity: World problems are solved together.
Small-scale ecology: World problems are solved locally.
The report is not based on scenarios but on an outline of the future, the "narrative." Unlike other
scenarios, possible future forms of society are outlined without asking how the development will pan
out in the next 50 years. These are alternative futures, so-called world views that are developed from
four combinations of two fundamental uncertainties in society's socio-cultural value model.
Uncertainties are issues such as deregulation, globalization, technological development,
centralization/decentralization, etcetera. The energy supply's physical possibilities and limitations and
economic relationships are rather seen as certainties.
Table 7.1 displays the assessment of the outlined world views, according to the Ministry of Economic
Affairs.

Free trade

Ecologic

Isolation

Solidarity

Security of supply
A . NA TIONA L P ERSP ECTIVE
B . CONSUM ERS P ERSP ECTIVE

Economic efficiency
C. COST A DJUDICA TION
D. JUSTIFIA B LE DIVIDING

Economic efficiency
E. EFFICIENCY
F. EM ISSIONS

+ = satisfied 0 = limited satisfied - = not satisfied

Table 7.1: Degree to which the energy supply in each world view meets today's criteria (Source:
Ministry of Economic Affairs, 2000)
ECN, Reference Projections for Energy and Emissions 2005 - 2020
In 2005, the Energy Research Centre of the Netherlands (ECN, 2005) used scenarios from the longterm outlook "Welfare and the environment 20022040" (WLO), published by CPB, RPB, MNP[2005].
Figure 7.4 displays a schematic overview of the scenarios concerned.

179

Part 4

International collaboration

Public responsibility

Personel responsibility

National sovereignty

Figure 7.4: Overview of the WLO scenarios (Source: ECN, 2005)


In the reference projection, ECN chose two WLO scenarios, namely Strong Europe (SE) and Global
Economy (GE), which are both strongly geared toward international collaboration. In SE, collaboration
is linked to public responsibility with a strong focus on the equal distribution of income, social security,
and a global climate policy in which the USA is also involved. In GE, there is a strong orientation
toward free trade but with little collaboration. Personal responsibility is the motto, and environmental
awareness is not articulated in strong legislation, which is why the international climate policy
ultimately fails.
ECN, The next 50 years: Four European Energy Futures
In its document The next 50 years: Four European Energy Futures ECN (2005) outlines four future
scenarios for Europe in which the key drivers are the relationship between oil demand and oil
production and the success or failure of the global climate policy. A global oil peak, reached when oil
production starts to decline permanently and demand can no longer be met, will lead to sharply rising
and permanently volatile oil prices. As a result, this would increase the pace of oil substitution
particularly in the chemical and transportation industries. The success or failure of the climate policy
will have a substantial impact on gas emission markets and the CO2 emission reduction potential of
future investments. Figure 7.5 displays the four scenarios for the European energy future.

180

Chapter 7 Scenarios for the Dutch Electricity Supply

Figure 7.5: Four European energy futures according to ECN (Source: ECN, 2005)
This results in the following contrasting storylines:
Firewalled Europe Oil production peaks in the period 2010 - 2020. No viable post-Kyoto
climate change policy emerges. The European energy sector turns back to coal and nuclear in
the next 50 years.
Fossil Trade Oil production follows oil demand smoothly in the period 2010 - 2020. No viable
post-Kyoto climate change policy emerges. The European energy sector continues business
as usual in the next 50 years.
Sustainable Trade Oil production peaks in the period 2010 2020. Post-Kyoto climate
policies develop effectively. The European energy sector turns to large-scale trade in
renewables in the next 50 years.
Fenceless Europe Oil production follows oil demand smoothly in the period 2010 2020.
Post Kyoto climate policies develop effectively. The European energy sector diversifies
strongly keeping all options open for the next 50 years.
European Union, ExxonMobil, BP, Ministry of Economic Affairs, Scenarios for Primary Energy Supply
World energy consumption has risen since the first oil crises but the energy import dependency of the
EU over this period decreased from 60% in 1973 to 50% in 1999. Because of the further increase in
consumption and the decrease in the EU's own production (see Figure 7.6) the overall energy import
dependence of the EU is likely to rise once again, reaching 70% within 20 to 30 years (European
Commission, 2001).This increase of import dependence can be seen for all forms of energy (see
Figure 7.7).

181

Part 4

Figure 7.6: EU-30 Energy production by fuel (Source: European Commission, 2001)

Figure 7.7: EU-30 Import dependence according to energy product (Source: European Commission,
2001)
Fossil fuels have to be imported over ever longer distances, usually from politically less stable regions
such as the Middle East, Russia, and Algeria. Moreover, the demand for fossil fuels may increase
strongly around the world (see Figure 7.8) by 1.7% per year (according to ExxonMobil projections).
The demand for oil and gas may account for about 60% of the total energy demand. Wind and solar
power are expected to grow by about 10% per year. However, the share of wind and solar power in
the total energy demand will only begin to approach 1% in 2030 (ExxonMobil, 2004).

182

Chapter 7 Scenarios for the Dutch Electricity Supply

Figure 7.8: Growth in primary energy sources (Source: ExxonMobil, 2004)


If we zoom in on the global electricity supply (see Figure 7.9), we see that production in Asia alone is
projected to grow by 40% as a result of rapid economic development. The right-hand graphic displays
the fuel inputs. The share of coal and gas is expected to rise to approximately 70% in 2030.

Figure 7.9: Growth in annual generation and fuel input (Source: ExxonMobil, 2004)

183

Part 4

The expected growth in electricity demand will put enormous pressure on the coal and gas markets in
particular. The demand for gas is particularly increasing in the USA and Asia. The limited number of
gas suppliers will have to invest heavily in large new gas projects. They want security of demand and
this results in long-term contracts. The European Union must have enough financially strong buyers
(such as GasTerra (formerly Gasunie Trade & Supply), Gas de France, E.ON Ruhrgas) that can stand
up to the few large suppliers and to strong buyers in other regions such as the USA and Asia (AER,
2005).

Billion m3

In the Netherlands, a good 50% of all of the EUs natural gas reserves is located (NAM, 2005).
However, with a cumulative production of 2,840 billion m at the end of 2005 and a proven reserve of
1,510 billion m at the beginning of 2006, only 35% of the initially estimated quantity is left (see Figure
7.10).

Remaining reserve

Year

Cummulative production

Figure 7.10: Natural gas reserves and cumulative production 19652005 (Source: Ministry of
Economic Affairs, 2006)
On January 1, 2005, the natural gas reserves were estimated to be 1,032 billion m for the Groningen
field, 149 billion m for the remaining territory, and 225 billion m for the continental shelf. In total, the
reserves amount to 1,510 billion m (Ministry of Economic Affairs, 2006). In 2005, the share of natural
gas from the smaller fields (onshore as well as offshore) comprised approximately 37 billion m of the
total amount, but that share is declining (NAM, 2005), which means that the Groningen field has to be
used more and more. In 2005, Gasunie Trade & Supply's annual gas sales were 80.4 billion m, 32.0
billion m of which were for the domestic market and 48.4 billion m for the export market, 45% of
which went to Germany (Gasunie Trade & Supply, 2006). The gross amount of natural gas produced
from the Dutch fields was 73 billion m. The rest was imported. Current sales figures mean that the
gas will be depleted within the next 20 to 25 years.
Based on the Gas Act for the Dutch production of natural gas, the Minister of Economic Affairs has set
a ceiling for the Groningen field in order to use the field as long as possible. For the period 2006 2015 this means a maximum of 425 billion m of natural gas from the Groningen field can be taken, an
average of 42.5 billion m (Ministry of Economic Affairs, 2006). The rest of the required natural gas for
the annual gas sales has to come from the smaller fields (is declining) and the import of natural gas;
see Figure 7.11.

184

Chapter 7 Scenarios for the Dutch Electricity Supply

Histo rical supply no n - Gro ningen

Histo rical supply Gro ningen

Supply co nfo rm explo iratio n plans

Expected supplies o ut still no t develo ped fields

Expected supply o ut still to disco ver fields

M aximum pro ductio n Gro ningen

Figure 7.11: Production of natural gas as of 1990, production prognosis for the smaller fields and the
maximum production space of the Groningen field (42.5 billion m / year (Source: Ministry of Economic
Affairs, 2006))
Import can be increased by pipelines from, for example, Russia, but also by building LNG terminals.
LNG increases market flexibility, meaning that it is very likely that by increasing the number of markets
for natural gas, the market that offers the most for the liquefied natural gas will determine the final
price in another gas market in order to obtain sufficient supply.
BP Statistical Review of World Energy June 2007 quotes for 2006 a reserves-to-production (R/P) ratio
of 40.5 years for oil, for coal, the R/P ratio is 147 years, and for gas approximately 63.3 years. Each of
these ratios is calculated by dividing remaining reserves at the end of the year by the production in
that year and is updated every year. The ratio is not a constant figure but changes if remaining
reserves and/or production rates (caused by growing demand) have changed. The oil reserves are in
the Middle East, the coal reserves in North America, Europe, Eurasia, and Asia Pacific, and the gas
reserves in the Middle East, Europe, and Eurasia. Although the worldwide demand for fossil fuels is
increasing, the reserves incorporated in the R/P ratio are still increasing too. For example, since the
beginning of the 1980s, the R/P ratio for natural gas has increased and has remained fairly constant
ever since (see Figure 7.12).

185

Part 4

Figure 7.12: R/P ratio for natural gas (Source: BP, 2007)
In summary, the information found in the referenced literature can be used to create the below matrix
of applicable driving forces and serves, in combination with scenarios for primary energy supply, as
input for the scenarios to be developed in Section 7.3.
Driving Forces

AER

Competition
Global climate policy
Market size
(Socio)economic
Sustainability/ Environment

TenneT

Ministry of
Economic
Affairs

ECN 1

ECN 2

X
X

X
X
X

Table 7.2: Matrix of driving forces from referenced literature

7.3

Developing Scenarios for the Model

7.3.1

Introduction

The description of the scenario development techniques, the objectives defined for the Dutch
electricity supply model, and the scenarios for the electricity and energy supply found in the literature
enable us to start developing scenarios for our model. First we will define the driving forces and
analyze why they are important for the model. Then, the driving forces' uncertainties will be assessed
and a matrix of extremes created. The selected and analyzed periods will then be used to create the
narratives or storylines for six possible scenarios.
7.3.2

Driving Forces and Uncertainties

The following step consists of determining the driving forces underlying the scenarios. Combining the
four driving forces mentioned in 7.2.2 (Wilkinson) with Table 7.2 results in:

186

Chapter 7 Scenarios for the Dutch Electricity Supply

Social dynamics Demographic issues leading to population growth or decline, prosperity.


Economic issues Macroeconomic trends and forces shaping the economy as a whole globally or
locally.
Technological development and diffusion New technologies, large or small energy
conversion techniques.
Environmental issues Environmental awakening or business as usual, boundless energy
consumption.
Fuel sources Security of supply, diversification of fuels used.
Political issues Electoral.
The driving force of social dynamics is dependent on, among other things, demographic and
prosperity issues. Demographic issues such as migration, aging, and fertility cause a population
growth or decline. A grow in case of a tolerant immigration policy and a decline in case of aging of the
population, less births and/or less fertility. The four WLO scenarios, ECN (2005) for example specifies
that the Dutch population will be between 16 and 19.5 million in 2040. The latter caused by a tolerant
immigration policy and a large number of births. Prosperity issues, specially the growth is one of the
main objectives of EU policy, and although competition between the other power houses (USA, Asia)
will continue to play a role, the EU will maintain its prosperity status. Increase in prosperity means that
more and more electrical appliances will be used and that, although the appliances are more efficient,
their use will probably still result in higher electricity consumption per head, see also Chapter 1.
As far as the economic issues are concerned, economy as a whole globally or regional oriented has a
direct influence on the gross domestic product and also electricity demand. Technology development
in a global economy, in which knowledge is shared to a large extent, leads to more technological
developments and improvements and has positive effects on investment levels and learning curves
(see Section 5.2.5). The direction in which the electricity supply will develop is not certain. Will the
current fairly strong regional orientation of the electricity supply be sustained or is there finally going to
emerge an EU-wide energy supply and demand market? This depends on a lot of factors like having a
EU-wide level playing field (Van Eck et al, 2004), capacity of the cross border connections etc.
One of the most important factors driving technology development is fuel-market prices and their
development. High oil prices and the ensuing high gas prices, for example, make the deployment of
and further research into alternative energy sources more attractive. It also stimulates exploration for
new fossil resources and improvement of conversion technologies. Such research could help improve
todays technology and boost the development of alternative conversion technologies. Another
uncertain factor is the further development of small-scale energy conversion as opposed to large-scale
conversion technology (see Part 3).
Environmental issues depend to a large degree on the regional or global climate policy pursued, the
price that buyers or rather voters, are prepared to pay for energy as such, and the awareness of the
consequences of excessive energy consumption, for example, for the greenhouse effect. Climate
policy with focus on energy savings leads also to an increase of the electricity share in the total energy
mix (Eurelectric, 2007).
Fuel sources and their development depend largely on the development of the remaining reserves in
combination with production rates required to fulfil the energy demand. In Figures 7.8 and 7.9 it can be
seen clearly that the demand for fuels as oil, gas and coal is further increasing whilst the remaining
reserves, specially for oil and gas are limited. It is unclear if additional reserves in addition to the easily
exploitable remaining oil and gas reserves can hold pace with increases in grow of energy demand.
There are however sufficient coal reserves, according to BPs Statistical Review the R/P ratio of coal
is enough for another 147 years of production based on 2006 figures. Another aspect and of influence
on security of supply is the place where remaining reserves are located. Most of the oil and gas are
located in politic unstable regions as the Middle east and Eurasia whilst the coal sources are located
in politic more stable regions as Europe, North America and Australia. More security of supply in the
future, less dependency on supplies from unstable regions and therefore reasonable fuel-market
prices leads inevitable to more diversification of fuels used with more focus on coal and uranium as
fuel source.

187

Part 4

Political electoral issues, primarily driven by social dynamics and economic issues, strongly
determine which policy is followed and what impact that policy will have on the energy supply. The
selected fuel diversification policy, climate policy (regionally or internationally driven), supervision
(regionally or internationally driven), and international collaboration strongly determine the energy
supply.
Although all of the driving forces carry a certain degree of uncertainty, we will have to make a choice.
In Table 7.2, the driving forces selected in four of the cases are economic issues and environmental
issues, whereby the latter can be considered to include global climate policy aspects. According to the
literature and the above description, the uncertainties in driving forces are primarily found in the
economic and environmental issues (see also Van Eck, 2007) and less in the social developments
and technological development and diffusion. There is however also a great deal of uncertainty in the
area of remaining fossil fuel sources, specially oil and gas in combination with increasing demands,
because of the EU's increasing external import dependency for its fuel supply. The selected driving
forces for the scenarios are displayed in the matrix below.
Regional
Oil/Gas peaks

Environmental

Non Environmental

Oil/gas does not peak


Global

Figure 7.13: Matrix of uncertainties


7.3.3

Horizons

To delineate a future scenario, we must define some horizons. We have tried to link them to the
milestones used in a variety of publications (EU, EZ, Vrom, Shell, RIVM).
Reference 2003
Although the year 2003 is not a scenario in itself, it is the starting point for developments that take
place in the scenarios developed below. The reference situation was described in detail in Part 1 in
relation to the supply and demand of electricity, primary fuel consumption, and CO2 emissions and will
also be used to validate the model (see Section 8.5).
2012
One of the EUs main goals is to boost the economy, create jobs, and stimulate competition and this is
an important reason why the energy market was liberalized (see also Chapter 1). The security of
supply is also important (European Commission, 2001) considering the EU's increasing dependence
on primary energy from other regions (Section 7.2.3). There are still enough fossil fuels for the coming
decennia, but it is uncertain if production and processing capacity can keep up with increasing
demand. The latter could drive prices up. Fuel diversification is an important instrument to decrease
the dependence on specific types of fuel.

188

Chapter 7 Scenarios for the Dutch Electricity Supply

As a consequence of the obligations that were entered into during the climate conference in Kyoto, the
EU has made further agreements (EU, 2002) to reduce the emission of CO2 and other greenhouse
gases. To contribute toward achieving the protocols European goals, it was agreed that the
Netherlands must reduce its greenhouse gas emissions by an average of 6% between 2008 and 2012
with respect to 1990, meaning an average emission level of 199 Mton CO2-eq. per year.
For the Dutch electricity production companies, this means that on average a total reduction of 5.8
Mton CO2/year must be achieved before 2012 (Essent, 2004).
Another indicative target set in EU Directive 2001/77/EG (EU, 2001) refers to the share of electricity
coming from renewable sources in the Netherlands. By 2010, this share must have increased to 9% of
the gross electricity consumption. For the then fifteen EU member states taken together, the indicative
target is 22%.
2025
Between 2012 and 2025 we are in the post Kyoto period. A new regime will be in force. As we expect
now, more stringent reductions will be in place for more countries. Also, USA, China and India will
become involved in emission reduction. Basically, EU objectives will have remained the same. The
dependence on primary energy from other regions will have increased because the EU's own reserves
will have decreased.
In the Netherlands, the government has formulated ambitious targets, a energy saving of 2% a year
and in 2020, 20% of the total energy consumption has to be from renewable sources and the emission
of greenhouse gasses has to be decreased by 30% compared with 1990 (Ministry of Economic Affairs,
2007). The government's plans are geared toward the realization of 6,000 MW of wind power in the
North Sea1.
2050 and Beyond
The energy conversion technologies chosen in 2025 are still relevant in 2050. The applied technology
is based on the same technical principles as in 2012 but will be more efficient and cleaner. The share
of renewable sources in the total electricity production is expected to have increased further. The
share of oil and gas in the total energy consumption however shall be limited caused by the declining
of remaining reserves in the world.
7.3.4

Narrative/Storyline Scenarios

The combination of driving forces, uncertainties, and the ensuing matrix result in the schema
displayed in Figure 7.14.

In the mean time , the European Commission has proposed certain targets for the post-Kyoto period until 2020
(January 23, 2008). The aim is to reduce green house gases by 20%, attain 20% renewables of final energy
consumption and to reduce energy demand by 20% compared to a certain baseline. Because these are still
proposals, we have stuck to our scenarios as defined above.

189

Part 4

Regional
Scenario B
Characteristics:
Lowest cost

Scenario A
Characteristics:
Ecological

Environment of

Sustainability

secondary importance

Self sufficient

Social dynamics

Self sufficient

Regional

Regional

Driving Forces

Economic issues
Model
Load growth

Technological

Fuel Mix

Environmental
Environmental issues

Fuel Cost
Conversion Technology

Non Environmental

Central / Decentral
Stimulation / CO 2 tax
Market / Regulation

Fuel sources

Political issues

Scenario D

Scenario C
Characteristics:
Ecological

Characteristics:
Lowest cost
Environment of

Sustainability

secondary importance

Collaboration

Collaboration

Global

Global

Global

Figure 7.14: Schematic overview of possible scenarios


The driving forces that are considered to be uncertainties economic issues and environmental
issues result in the extremes global/regional and environmental/non environmental which are the
basis on which the scenarios are developed. These extremes are also found in Energy and Society in
2050 (Ministry of Economic Affairs, 2000) and TenneT's capacity plan (2002). In the Ministry of
Economic Affairs explanatory memorandum, the extremes Global and Regional refer mainly to the
national and international economy and the consequences for the energy supply. In TenneT's capacity
plan, the extremes Global and Regional refer mainly to an open electricity market as opposed to a
protectionist approach. In Reference Projections for Energy and Emissions 2005 - 2020, ECN
focuses on international collaboration with the extremes much attention or little attention to
environmental awareness. In The next 50 years: Four European Energy Futures, ECN includes the
uncertainty of oil supply in the case of the presence or absence of an international climate policy. The
AER (2004) outlines the consequences of consolidation on the electricity markets, and reaches the
extremes open electricity market and protectionist approach with less emphasis on use of
renewable energy and lowest costs than is demonstrated in the Ministry of Economic Affairs'
explanatory memorandum and TenneT's capacity plan.
In this thesis, we obviously focus on the extremes mentioned. However, much attention, is also paid to
the impact that the selected electric supply configuration will have on the security of supply, on
economic aspects, and on environmental aspects. The security of the fuel supply will also be
highlighted.
Four scenarios for the Dutch electricity supply that will be quantitatively calculated with the model are
described below.
Scenario A: Environmental within National/Regional Borders with Emphasis on Renewable Sources
In this scenario, the development of renewable sources has a high priority in particular with regard to
the exhaustibility of fossil fuels, the desired far-reaching decrease in CO2 emissions, both nationally
and internationally, and the desired transition to an entirely CO2 emissions free energy supply. In this
transition phase, natural gas is the fuel of preference. The national gas reserves are handled with care
by phasing out existing export contracts so that until about 2050 enough natural gas is available for
domestic consumption. Natural gas is used where possible in high-quality CHP applications as well as
for indoor heating and hot tap water in areas without district heating. Instead of building more nuclear

190

Chapter 7 Scenarios for the Dutch Electricity Supply

or coal power plants, large biomass plants are built in which biomass from local resources is used,
stimulated with government measures. CO2 capture and storage systems are introduced as of 2020
on a modest scale in multi fuel conventional and IGCC units. Although energy use per head is
decreasing, the share of electricity in the total energy consumption continues to rise. After much public
debate it is decided to develop large offshore wind farms with the ultimate goal of generating 6,000
MWe at sea and 2,000 MWe on land by 2020. Measures from government are taken to support this
large scale wind development. By then, however, it will still not be possible to store electricity on a
large scale meaning that the generation of wind power will have to be integrated into the total
electricity supply system in the best possible way. In this system, large-scale CCGT power plants are
required to compensate for the winds variability, because of their excellent regulation curve between
part and full load operation and their short startup time. The variability of the wind means however that
these CCGT power plants are not the ideal candidate to supply heat, which is why it is not considered
for heat delivery application.
The development of one big open European electricity market fails as a result of protectionist
tendencies. The markets remain regionally organized. Hardly any electricity is transported via the
interconnections except for that portion that is necessary for emergency situations in which countries
continue to help each other.
Scenario B: Non Environmental, Pursuance of Lowest Costs within National/Regional Borders
In this scenario, monetary profit predominates. Everything, including electricity, must be generated at
minimum cost so that profits can be maximized. Developments in renewable energy have come to a
complete standstill and investments are made in CHP only in places where a fast return on investment
is guaranteed. The domestic natural gas reserves are quickly depleted due to continuing export and,
in combination with expensive import, are just sufficient to cover the domestic gas needs for indoor
heating and hot tap water in areas without district heating. Because coal and uranium are sufficiently
available on the world market for a relatively low price, the large-scale use of coal and nuclearpowered electricity generation is preferred. Energy use per head is increasing and electricity use in the
total energy consumption is increasing even more. The development of one big open European
electricity market fails as a result of protectionist tendencies. The markets remain regionally organized.
Hardly any electricity is transported via the interconnections except for that portion that is necessary
for emergency situations in which countries continue to help each other.
Scenario C: Environmental issues are Optimized Around the World with an Emphasis on Renewable
Sources
In this scenario, economic growth is high as a result, among other things, of far-reaching international
collaboration and economic globalization. Because the fossil fuel reserves are constantly decreasing,
combined efforts are being made to find global solutions to the energy problem so that economic
growth is not hindered. Natural gas is used where possible in high-quality CHP applications as well as
for indoor heating and hot tap water in areas without district heating. Innovative solutions that focus
primarily on renewable energy and its large-scale storage are being developed and stimulated with
government measures. These solutions will enable supply and demand to be disconnected and fit
renewable energy better in the overall supply system. The desired transition to an entirely CO2
emissions free electricity supply in combination with international collaboration on development of new
technology has made the introduction CO2 capture and storage systems possible as of 2020 on a
large scale in multi fuel conventional and IGCC units. A large share of biomass from international
sources is applied in the fuel mix of this power plants and is stimulated with government measures.
The energy use per head is decreasing but the share of electricity in the total energy consumption
continues to rise in spite of increasingly efficient processes.
One big European electricity market is steadily developing, and because of the far-reaching integration
of the energy markets they come close to a copperplate situation in which a level playing field is
created. The cost of electricity transportation weigh heavily in the decision where to build new
generation capacity. The inter-connector capacity is expanded modest, and the Netherlands plays an
important role in the European electricity supply because of its favorable location for electricity
production.

191

Part 4

Scenario D: Non Environmental, Pursuance of Lowest Costs in an Internationally Oriented Energy


Supply
In this scenario, economic growth is high as a result, among other things, of far-reaching international
collaboration and economic globalization. Fossil fuel reserves are still sufficiently and largely available
mainly as a result of new findings, which have increased the remaining reserves constantly, and the
installation of additional production capacity. The energy prices are low and therefore the consumption
of mainly electricity relatively high. Developments in renewable energy have come to a complete
standstill as a result of short-term financial profit, which does not accommodate for renewable energy.
CHP is used in places with a high return on investment. The same applies to other new conversion
technologies.
One big European electricity market is steadily developing, and because of the far-reaching integration
of the energy markets they come close to a copperplate situation in which a level playing field is
created. The cost of electricity transportation in the European system weigh heavily in the decision
where to build new generation capacity, which is why the inter-connector capacity is not expanded,
and the Netherlands plays an important role in the European electricity supply because of its favorable
location for electricity production.
Each of the above described scenarios has its own requirements for the type of fuel supply and use of
new technologies, these starting points for the scenarios are given in Table 7.3:
Scenario

Gas (land)

LNG

Coal

Nuclear

Biomass

Wind

Yes

No

Yes (limited)

Yes, local
resources

Yes, without
storage

Yes (limited)

No

Gasification
coal with
biomass with
modest CO2
capture &
storage
Yes

Yes

No

No

Yes

Yes

Yes (limited)

Yes,
international
resources

Yes, with
storage

Yes

Yes

Gasification
coal with
biomass and
CO2 capture
& storage
Yes

Yes

No

No

Table 7.3: Starting points for the fuel supply in the scenarios
As can be noted in Table 7.3 the technology developments in scenario C are more far ahead
compared to scenario A and this is caused by the international and joined efforts for developments of
renewable energy sources including storage of wind energy and CO2 capture and storage.
7.4

Summary

Scenarios for the Dutch electricity supply have been developed. First, the driving forces have been
determined, and then their uncertainties which have been used to create a matrix of extremes.
Combined with definitions of their horizons, four possible future scenarios have been developed.
These scenario cover trends in supply and demand, protectionism versus globalization, and the
importance of the environment. The four scenarios serve as input for the simulation model.

192

Chapter 8 Dutch Electricity Supply Model

Chapter 8
8

Dutch Electricity Supply Model


8.1

Introduction

The Dutch electricity supply model that was developed for this study is a long term technical economic simulation model for the energy demand and supply system. The model combines the
existing technical energy supply system (Chapter 4) with possible future technical energy supply
systems (Chapter 6) based on four potential societal scenarios (Chapter 7). Main purpose of the
model is to analyze the four scenarios with respect to the balance between the environment,
economics and security of supply. In Figure 8.1 an overview of the models place in this thesis is given.
Influences
Surrounding

Part 4 & 5

Technical System
(Part 2 & 3)

Scenario's

Input

Main m odel
Balance Dem and - Supply

Output

Conclusions

Input

Figure 8.1: Schematic overview of the model's place in this study


The scenarios developed in Chapter 7 and the technical system constitute the input for the electricity
supply model of which the design will be discussed in the next sections. After the objectives for the
model have been formulated and an inventory of the models found in the literature has been made,
the model's starting points and input/output are defined. The mathematical/economic structure of the
model is then described and the model is validated. The model's input and analysis of the results for
the developed scenarios is described in detail in Chapters 9 & 10.
8.2

Objectives, Delineation, and Simplification of the Model

The model's objective can be described as follows:


Use a combination of technical system input, applicable external factors and economic modules to
make calculations that result in quantitative outcomes for the developed scenario.
The quantitative outcomes are used to draw conclusions about the balance between the environment,
economics and security of supply for the different scenarios and compare them as described in the
introduction. The model however is based on a transparent and manageable techno-economic world,
not affected by the politico-economic value chain of the energy supply. In reality factors like debate on
public versus private interest, the position and behavior of stakeholders, the market structure, the
(environmental) regulation have influence (see Van Eck, 2007).
The model is not a market model that describes year to year, month to month or day to day market
behavior rather it is primarily developed to calculate the long-term effects (scenarios) using the current
electricity supply and demand system with its technical input and economic preconditions and their
expected development in the future under the different scenarios.

193

Part 4

Delineation and simplification of the model:


General
The year 2003 is the starting point from which the electricity supply and demand options
(including import) for the chosen scenarios are determined (see Chapter 2 en 3).
The year 2003 is the reference year, but the model uses the progress of the curve for 2001
(8760 h, starting on Monday, no leap year) for the input and output because the dispatch
module PowrSym3 uses 2001 as an internal calendar (see Section 8.4.1).
The model calculates the balance between electricity supply and demand for every hour for
the Dutch supply as a whole and for one year (forecast) rather than for individual market
players (excluding decentralized electricity demand).
Effects of daily imbalance as a result of (small) differences in supply and demand are not
taken into account.
Electricity Demand
The shape of the electricity load curve is derived once using TenneT data from 2005. The
progress of the TenneT curve is converted to the year 2001 corresponding with the internal
calendar of PowrSym3. Own research is conducted to fill in the missing data and obtain a load
curve for the Netherlands, which is scaled back to the electricity demand of 2003.
Increase of the electricity load curve as a consequence of economic development in the
scenarios is calculated by applying a growth percentage equally for every hour in the year on
the original TenneT curve.
The load shift between peak and off-peak periods can be determined by adding a percentage
to or deducting a percentage from the peak and off-peak hours in the original curve.
Heat Demand
The shape of the district heat load curves is based on real heat load curves of three larger
district heating areas in the Netherlands.
Where possible, own analysis was carried out to calculate the expected/possible increase in
district heat demand for the areas where district heating is or can be made available and
which have upward potential.
The impact of global warming on the number of degree-days for future heat demand is not
taken into account.
The effects of using heat storage buffers are not taken into account.
The shape of the process heat load curves are assumed to be a flat line.
The additional need for cooling as a result of global warming is assumed to have been
deducted from the increase in electricity demand and is not taken into account.
Electricity Supply
The effects of limiting the supply power (cooling water limitations, reduction in gas turbine
capacity) as a result of ambient temperature is not taken into account.
The reliability and availability of the units is specified using an assumed percentage which is
supposed to remain constant throughout the whole period of simulation.
The effects of using storage facilities for electricity are partly taken into account.
Emissions
NOx emission is not calculated because it is expected to further decrease as a result of a
legally imposed NOX emission trading system and the effects of the IPPC which lead to the
implementation of improved technologies for DeNOx at coal units and Dry Low NOX at gas
turbines.
SOx emission is not calculated.
Grids

194

Limitations on the national grids are not taken into account.


For the inter-connectors, the normal securely available transport capacity is assumed.
Limitations are not taken into account.

Chapter 8 Dutch Electricity Supply Model

8.3

Structure of the Model

8.3.1

Introduction

The model's design is explained in this section. The design uses today's Dutch electricity supply
system: the electricity and (if applicable) the heat from electricity production companies, other
electricity producers Chapter 4) and import and export. It calculates the total fuel consumption (TJ)
and the environmental load (CO2), the total cost of the electricity and heat generation and the security
of supply i.e. the available reserve capacity. Other studies found in the literature for the Dutch situation
focus often on sub sectors alone such as developments in CHP or renewable energy (wind) only. This
model simulates all of the links in the electricity supply chain in order to shed light on the effects of
choices made in sub sections as well as environmental load, the related costs, and the security of
supply.
8.3.2

Short review of energy models, comparison with our model

Models have become standard tools in energy planning. Energy models are developed for efficient
energy planning, forecasting and optimization of energy sources (Jeberaj et al, 2004). The various
types of energy modeling which have been reviewed and presented by Jeberaj are:
Energy planning models, for centralized and decentralized planning. Decentralized energy
planning (DEP) is a concept of recent origin with limited applications and literature shows that
different models are being developed and used worldwide (Hiremath et al, 2005).
Energy supply-demand models, for estimation of fuel resources, energy demand forecasting to
project energy requirements, analyzing electricity demand etc.
Forecasting models, have been formulated using different variables such as population,
income, price, growth factors and technology and are categorized into two groups, commercial
energy models and renewable energy models which can be split-up in solar, wind and biomass
and bio-energy models.
Optimization models for optimization of fuel energy balance, minimizing capital and operation
cost of energy supply, dispatch models to optimize the use of generation units over time and
determine the resulting market prices based on the units merit order (the last unit determines
the final balance price) etc.
Energy models based on neural networks, intelligent solutions, based on artificial intelligence
(AI) technologies to solve complicated practical problems.
Emission reduction models, for developing CO2 emission scenarios.
Besides these energy models economic models are developed for the electricity sector mainly geared
toward analyzing the rational market and the irrational market. The rational market is a well-functioning
market that is characterized by transparent supply and demand price levels that are explainable
through analysis and the irrational market is characterized by a lack of transparency, situations in
which extreme price peaks occur, for example, on the APX (Lapuerta et al, 2001), or in California (Kip,
2001). A variety of models is developed to calculate the competition on the electricity market and to
analyze market power (Boisseleau, 2004).
As stated in Section 8.2, our model is not a market model that describes market behavior; rather it is
primarily developed to calculate the long-term effects (scenarios) for the balance between the
environment, economics and security of supply. It is assumed that trade in energy takes place in
perfect competition in a rational market, existing in the future as well, with completely transparent (cost)
price levels and is not negatively affected by the politico-economic value chain of the energy supply at
all. In relation with the various types of energy modeling as described above, our model is an energy
supply-demand model with forecasting and optimization sub-models.
In the literature two relevant PhD studies were found which can be compared with this thesis. The first
study was carried out at the University of Leuven (K.U. Leuven, Belgium), the second at the University
of Groningen.

195

Part 4

The first study The modeling of large electricity-generation systems with applications in emissionreduction scenarios and electricity trade Voorspools (2004) a methodology has been developed to
simulate electricity generation and trade in a set of interconnected zones. The resulting model can be
used for impact assessment of emission reduction scenarios and electricity trade.
The second study Interactive simulation of electricity demand and production of Benders (1996) the
focus is on development and application of a set of tools for the interactive exploration of feasible
(mid- to long-term) electric power system futures in term of their technological, socio-economics and
environmental impacts.
For the models briefly described above and the applied model in this study the differences in choices
made can be pointed out and deviations from our model can be determined, see Table 8.1, without
giving a judgment about the models themselves.
Driving parameters for electricity demand in sectors
Future Scenarios
Electricity demand curve on hourly bases
Heat demand curve on hourly bases
Simulation on hourly basis
Detailed input parameters large power plants
Detailed input parameters individual small scale plants
Handling of combined Heat/Power generation
Renewables supply
Storage electricity
Restrictions and cost of transmission grids
Unit Commitment, Economic Dispatch
Environmental gaseous emissions
Environmental solid waste and particle emissions
Annual cost of generation
Security of Supply or Loss Of Load Probability

Voorspools
NA
A
A
NK
A
A
A
A
A
A
A
A
A (CO2 only)
NA
A
A

Benders
A
A
NA
NA
NA
NA
NA
NA
A
NA
NA
NA
A
A
A
A

Our Model
NA
A
A
A
A
A
NA
A
A
A
NA
A
A (CO2 only)
NA
A
A

Table 8.1 Choices made for the three compared models, A = Applicable, NA = Not Applicable, NK =
Not Known
With the results of Table 8.1 it can be concluded that the model of Voorspools and our model have
been build on several equal starting points with respect to the input for supply and demand and the
used optimization routines. The models are, however, different with respect to the purpose of the
model. With the model of Voorspools electricity production and the trade in electricity can be simulated
for interconnected zones. His model can be used to calculate what influence emission reduction
objectives have on electricity production and trade in electricity for the interconnected zones. Our
model is primarily developed to simulated the long-term effects for the balance between environment,
economics and security of supply for different chosen scenarios. With our model, we can find out
which scenario suits best for emission reduction and against what cost en consequences for the
security of supply. Our model is also capable in simulating transports between interconnected zones
but this is limited in this thesis caused by choices made for the scenarios. Finally, the model of
Benders was developed for educational as well as for scenario studies. The model has limitations
because the production module is based on the load-duration curve and therefore no chronological
information is available. Due to the period it was built the model was also not occupied for the large
introduction of co-generation in the mid nineties which for the Netherlands growth to an installed
capacity of almost 40% as of 2003.

196

Chapter 8 Dutch Electricity Supply Model

8.3.3

Model Structure and Sub-model Description

The model's structure is displayed in Figure 8.2. The main model includes the demand curves for
electricity and heat demand (see Section 8.3.5) required for the calculations. The supply part is
divided into 3 sub-modules, the electricity production companies, other electricity producers and
import / export. The sub-module of the electricity production companies is connected with a database
containing the technical data of existing and new to build power plants required for optimization
calculations in the Unit commitment/Dispatch routine (see Section 8.4.1). The sub-module of the other
electricity producers consists of sub modules for industrial CHP, DH CHP, Small CHP, Non Fossil and
Waste Incinerators. These modules incorporate a large number of existing and future technologies
characterized by a number of technical data such as base- and part load efficiency, specific emission
factors, production relations between heat and power and other data. A very large part of process- and
district heating is supplied by CHP and small CHP which gives a small adjustable band for the
production of electricity and an interdependence with electricity production of the electricity production
companies. For Waste Incinerators, production of electricity is mainly driven by the availability of
waste and Non fossil electricity production depends completely on the availability of non fossil supply
sources like wind and sun (see Section 8.4.2). Finally, the sub module of Import/Export consists of
import or export profiles on the existing or future cross border connections (see Section 8.4.3). The
results of the sub modules are incorporated in the main model and the total results can be displayed.
Economic Input

Fuel

Subsidy

Demand Curves

CO2

TECHNICAL DATABASE

ECONOMIC STRUCTURE

Electricity production com panies

Unitcom m itm ent / Dispatch

IND CHP > 250 MWe


IND CHP < 250 MWe
DH CHP

DEMAND CURVES

Emissions of CO2
Annual system cost
Reserve factor

Output

Other electricity producers

Main model
Balance Demand - Supply

Technical system Input

SMALL CHP
NON FOSSIL
WASTE INC

Im port / Export

Figure 8.2: Structure of the model


The modules in the model are designed in the same way as the electricity balance used by Statistics
Netherlands (CBS) so that the results of the reference case for 2003 can be validated and the future
scenarios can be compared with the base case. For 2003, part of the electricity balance of Statistics
Netherlands is given in Table 8.2 for the Central and Decentralized Industrial sector. The Central
sector in Table 8.2 is named Electricity production companies in Figure 8.2 and the Decentralized
sector in Table 8.2 is named Other electricity producers in Figure 8.2. In this thesis, decentralized
generation refers to electricity generation and consumption on an industrial or other location it self,
with other words this electricity is not supplied to the public grid.

197

Part 4

Production facilities electricity


Subjects

Production electricity
Physical units
Production electricity Production steam/heat

MWh
Periods
2003
2003
2003
2003
2003
2003
2003
2003
2003
2003
2003
Not CHP
2003
2003
2003
2003
2003
2003
2003
2003
2003
2003
Decentralized
CHP
2003
2003
2003
2003
2003
2003
2003
2003
2003
2003
Not CHP
2003
2003
2003
2003
2003
2003
2003
2003
2003
Centraal Bureau voor de Statistiek, Voorburg/Heerlen 2007-09-28

Industrial sector

CHP

Typ of installation

Central

CHP

Total installations
Gas engine
Steam turbine
Ccgt-unit
Gas turbine
Nuclear unit
Hydro power plant
Wind turbine
Solar energy
Other installations
Total installations
Gas engine
Steam turbine
Ccgt-unit
Gas turbine
Nuclear unit
Hydro power plant
Wind turbine
Solar energy
Other installations
Total installations
Gas engine
Steam turbine
Ccgt-unit
Gas turbine
Nuclear unit
Hydro power plant
Wind turbine
Solar energy
Other installations
Total installations
Gas engine
Steam turbine
Ccgt-unit
Gas turbine
Nuclear unit
Hydro power plant
Wind turbine
Solar energy
Other installations

Installed capacity

TJ

Electrical capacity

Thermal capacity

Number of units

MWe

MJ/h

absolute

22633730

34596

4476

11228782

20
-

10207665

4828

1944

2322000

12164183

27921

2440

8082612

14

261882

1847

92

824170

43890448

9693

23

29310293

7302

11

10516747

1827

45419

115

4017989

449

26612618

176900

5080

46411276

3519

5120467

26334

1494

7949475

3354

1547340

34648

368

10544269

38

14054390

60064

2269

16415445

39

5829371

55707

929

11420566

75

61050

146

21

81521

13

5003509

1706

1659

1962492

1674

24906

20

57954

33

3382764

1706

601

1734938

27

169600

72023

37

1318000

906

1595

30740

46

175076

46

.
-

12

Table 8.2: CBS electricity balance for reference case in 2003 (Source: http://statline.cbs.nl)
CBS's definition of CHP is based on the supply of heat, which means that the three coal-fired power
plants with heat supply and with a capacity of 1,944 MWe fall under CHP within the Central Industrial
sector, meaning that CBS estimated the total installed CHP capacity at approximately 9,556 MWe in
2003. This definition differs from this thesis in which the definition of CHP excludes this capacity and
the installed CHP is hence estimated to be approximately 7,829 MWe (see Table 3.2 of Chapter 3).
8.3.4

Model Input and Output

With the model input parameters all kind of scenarios can be simulated. With the model output the
balance between the environment, economics and security of supply for the different scenarios can be
analyzed and compared and conclusions can be given. The input and output is described in the next
part.
Model Input
Demand Curves
1. Annual pattern of the total Dutch electricity load displayed per hour (see Section 8.3.5).
2. Annual pattern of process steam demand in industrial projects of CHP of the electricity production
companies is assumed to be constant. Annual pattern of heat load for DH projects of CHP of the
electricity production companies varies with the seasons and day and night pattern. The heat load
is displayed per hour.
3. Annual pattern heat load for DH projects as a function of predicted future degree-days and steam
demand in industry projects, assumed to be constant in the future.
4. The heat load for process steam and district heat for CHP and small CHP of the other electricity
producers, is displayed per month.

198

Chapter 8 Dutch Electricity Supply Model

Economics
1. Fuel price scenarios for natural gas, coal, light oil and uranium.
2. CO2 price scenarios as surcharge on the fuel price.
3. Stimulation measures (a fee that is added to the electricity market price) for the use of biomass
fuel and/or renewable electricity generation.
Technical System
1. The electricity production companies generation plants with heat consumption curves, minimum
and maximum electrical load, minimum up- and down time, technical preconditions, and heat
supply obligations.
2. The other electricity producers generation plants displayed in clusters with heat consumption,
electrical and thermal output, minimum and maximum load, and heat supply obligations.
3. Exploitation costs of the different types of electricity generation as startup cost and variable and
fixed operation and maintenance cost for electricity production companies generation plants, and
fixed operation and maintenance cost for other electricity producers generation plants. For
generation plants running on natural gas, the gas capacity cost.
4. Investment cost of the different type of nowadays applied and future electricity generation
technology.
Maintenance
1. Monte Carlo-based maintenance schedule (planned) and forced outages as a result of
disruptions (unplanned) of the electricity production companies generation plants. Percentages
for planned maintenance and forced outage are forecasted for every unit.
2. Average availability factor for each month based on maintenance schedule (planned) and failures
as a result of disruptions (unplanned) of the other electricity producers generation plants.
Model Output
For the electricity generated by the electricity production companies individual generation plants and
the electricity generated by the other electricity producers clustered production capacity the following
output results are given separate and combined:
1. The total amount of electricity generated
2. The total amount of heat generated.
3. The total amount of fuel required.
4. The total CO2 emissions and CO2 emissions per MWh.
5. The total operation cost and operation cost per MWh. Operation cost are the fuel cost minus the
fuel based revenues of the delivered process and/or district heat.
6. The total dispatch cost and dispatch cost per MWh. Dispatch cost are the operation cost plus
variable exploitation cost as CO2 cost, startup cost and variable operation and maintenance cost.
7. The total dispatch plus fixed exploitation cost and dispatch plus fixed exploitation cost per MWh.
Fixed exploitation cost are the fixed operation and maintenance cost and gas capacity cost.
8. The total integral cost and integral cost per MWh. Integral cost are the dispatch plus fixed
exploitation cost plus the investment cost leveled (depreciation, interest) according to the lifetime
of the technologies.
9. The reserve factor defined as total installed power divided by peak load. Three reserve factors are
calculated according to TenneTs method (2006), (1) without import capacity, renewable capacity
count for 100%, (2) without import capacity, renewable capacity count for 20% and (3) with import
capacity, renewable capacity count for 20%.
10. Reserve power, the total installed power minus the peak load.
8.3.5

Demand and Supply Balance

As already mentioned, electricity is a product that cannot be stored in large quantities, meaning that it
has to be supplied when it is requested. Changes in the electricity demand are absorbed by switching
on, increasing the electric power supplied, decreasing the power, and switching off units to create a
balance between supply and demand. Before the Electricity Act of 1998 came into effect, this was
centrally regulated by the Association of Energy Producers with a central dispatch. In the present day
electricity supply system, this is the responsibility of the Transmission System Operator (TSO) TenneT

199

Part 4

and a system of Program Responsible (PR) was introduced. With this system, the energy programs of
individual parties (operating on the electricity market) who are program responsible are approved on
the day of preparation, in order to secure the balance between demand and supply on the day itself
(Tennet, 2002). This has led to the organization structure displayed in Figure 8.3.

Figure 8.3: Organization of the Dutch electricity supply (Source: TenneT, 2002)
The PRs electricity grid supply programs are sent to the local grid operators and the umbrella grid
operator TenneT twelve hours in advance. TenneT continuously monitors the balance between supply
and demand in the Netherlands to ensure that trans-border exchange occurs according to the agreed
programs. TenneT corrects a possible imbalance by deploying the imbalance capacity contracted at
generation plants.
In principle, each PR with own generation capacity, has capacity that it can use flexibly to balance his
own supply and demand discrepancies and prevent an imbalance occurring elsewhere that can be
expensive to compensate.
Electricity production companies and other electricity producers (with generation plants with an
installed capacity larger then 10 MWe), have to provide TenneT with their generation data (DTE,
2002). With this data TenneT can centrally determine how big the reserve capacity is and adjust
supply and demand in the event of a disruption.
All the electricity production companies are currently Program Responsible. The other electricity
producers (companies with own industrial CHP generation, horticulture companies with small CHP,
independent wind energy producers etc.) often let energy companies carry out their PR tasks for
which a handlings fee has to be paid.
As far as the larger PRs and the system in general are concerned, the procedure to fulfill the demand
and supply balance is as follows:
One year before
The trade department of the electricity production company has signed long-term or annual
contracts with buyers such as industries, large and medium-sized companies, clustered
consumers etc. based on forward electricity and fuel prices.
All of the electricity which has to be supplied via the public grid for the whole year is displayed
in the public grid supply curve of the production company.
The annual statements of other electricity producers who have sold (part) of their generated
electricity to the electricity production company of interest, are processed in the supply curve

200

Chapter 8 Dutch Electricity Supply Model

by deducting them from the public grid supply curve. This electricity is the expected part of the
generated electricity of industrial CHP, district heating CHP, small CHP, non-fossil and waste
incinerators to feed back into the public grid. These supplies are based, among other things,
on forward electricity and gas prices (see Section 8.4.2).The decentralized generation has
already been deducted from the generated electricity.
Mandatory import contracts are processed in the supply curve by deducting them from the
public grid supply curve.
An annual schedule is created for the expected commitment of the electricity production
company own generation plants.

One week before


The supply curve is updated based on weather forecasts or other events that are scheduled to
take place that week.
The supply that other electricity producers are expected to feed back into the public grid is
updated based on the weekly schedules submitted by the producers.
Import contracts (annual and weekly contracts) are processed in the supply curve.
A weekly schedule is created for the expected commitment of own generation plants.
The day before
The supply curve is updated based on weather forecasts or other events that are scheduled to
take place that day. Discrepancies lead to a shortage or a surplus.
If changes are made, the supply that other electricity producers are expected to feed back into
the public grid, is updated. The short-term wind forecast in particular is important for the
expected production of wind power.
Import contracts (annual and weekly contracts) are included into the supply curve.
A number of economic considerations are weighed: commit own generation plants, buy from
the national or international power exchange markets, or sell on the national or international
power exchange markets (refer to Boisseleau, 2004, for an extensive description of power
exchange markets)
The day itself
TenneT controls the supply and demand discrepancies that occur at the production companies
by deploying contracted regulating power. TenneT also checks which PR was responsible for
the imbalance and is hence responsible for the associated costs.
The PR uses economic considerations (expected imbalance prices) to decide if it will commit
its own generation plants to control itself the supply and demand discrepancy.
Model
For the models demand and supply balance we selected an approach in which a whole year is
calculated on an hour by hour basis. We use previously accepted starting points for the use of supply
of other electricity producers (see Section 8.3.3). Import/export are fixed power blocks for peak and
off-peak periods. We must remember that the model is used to calculate scenarios, not markets or
submarkets. The following process, which is as close to the current situation on the electricity supply
market as possible, was selected (see also Figure 8.4):

The trade departments of the electricity production companies have, based on forward
electricity and fuel prices, signed contracts with buyers such as industries, large and mediumsized companies, clustered consumers etc. The models calculations are based on assumed
annual supplies.
The Dutch electricity supply via the public grid for one year is displayed in a here after called
supply curve.
Electricity supplied by other electricity producers via de public grid is deducted from supply
curve (Dutch electricity supply via the public grid plus decentralized generation is the total
Dutch electricity supply!)
Import contracts are processed in the public grid supply curve by deducting them from the
total supply curve.

201

Part 4

The minimum production capacity is determined that the heat supplying generation plants and
the generation plants which burn blast furnace have to generate in order to fulfill their heat
delivery and blast furnace gas obligations.
The unit commitment/dispatch module determines the commitment of the generation plants
(see Section 8.4.1).

IND
CHP

DH
CHP

GE
CHP

NF

WI

PRODUCTION

IMPORT

LOAD 1

+/+/+

DECENTRALIZED

LOAD 2 = LOAD NETH TOTAL - LOAD 1

PUBLIC GRID (including Grid Loss)

LOAD NETHERLANDS TOTAL

E - DEMAND

Large Consumers

Small Consumers

GRID LOSS

Export

Figure 8.4: The balance between supply and demand applied in the model
Heat Supply
Heat supply obligations of generation plants for heat supply to an industrial process or district heating
have influence on the supply balance as treated before. Forecast is of importance because in most
generation options heat supply leads to a lower possible maximum and minimum electrical load of a
unit (see Section 8.4.2). Heat supply to an industrial process often takes the form of continuous steam
supply throughout the year and maximum and minimum possible load of the unit is more or less fixed
and easy to forecast. Heat supply for district heating, on the other hand, is extreme sensitivity to ambient
temperature and consumption patterns and can trigger strong seasonal and daily fluctuations in the
demand for heat. Forecast is therefore of importance, however, less complicated compared with the
supply of electricity because district heat can be stored to a limited degree (buffers, grid buffering).
For heat supplying units of the electricity production companies yearly, weekly and day before
planning are made for the predicted heat demand and minimum and maximum possible load of the
units is calculated as input for the unit commitment/dispatch module which determines the
commitment of the generation plants.
For heat supplying units of the other electricity producers the predicted heat demand is taken into
account for the predicted electricity supplied via de public grid.
In the model, the heat demand input differs for the electricity production companies and other
electricity producers. For the heat supplying units of the electricity production companies, annual heat
load forecast curves are made on an hourly basis. The shape of the forecast curve for heat supply to
an industrial process is a constant. The shape of the forecast curve for district heat demand is based
on real district heat (production) curves (see Figure 4.15 of Chapter 4) in witch the total district heat
demand is a function of the standard degree-days year. The annual heat load forecast curves for
supplying units of the other electricity producers are modulated as a block (see Section 8.4.2) and
forecasts are made on a monthly basis.
.

202

Chapter 8 Dutch Electricity Supply Model

8.4

Mathematical/Economic Structure of the Model

8.4.1

Economics of Electricity Generation Companies, Unit Commitment/Dispatch

In Section 8.3.3 the model's structure has been given and the 3 sub-modules of the supply part were
described briefly. In this section, the economics of the sub-module for electricity production companies
is described in more detail. The sub-module contains a database containing the technical data of
existing and new to build power plants. The problem we face is the economic load division of the grid
coupled individual generation units with the applicable constraints in such a manner that the demand
load curve can be fulfilled at all times and with a minimum of generation cost.
Suppose we have three units, a conventional coal-, a hot wind box CCGT- and a CCGT unit. The
relation between the required heat (=fuel) energy and the load for this units was already given in the
specific heat consumption curves for various conventional units in Section 4.7. This curves can be
transferred into curves with the total heat consumption against the load, see Figure 8.5.
4,000

3,500

Total heat consumption [GJ/h]

3,000

2,500
Conv Coal
2,000

HW CCGT
CCGT

1,500

1,000

500

0
100

150

200

250

300

350

400

450

Load [MWe]

Figure 8.5 Total heat consumption curves


The curves can be mathematically expressed with a second order polynomial function:

F ( x) = a + bP + cP 2
where:
F(x)
a
b
c

Total heat consumption (in GJ/h)


Fixed heat use coefficient (fixed offset)
Specific heat use coefficient (steepness of curve)
Specific heat use coefficient ( bend of curve)

The first order derived of the previous function leads to:

dF ( x)
= b + 2cP
dP
The results are expressed in Figure 8.6.

203

Part 4

10

d (Heat Consumption (GJ/h))/d (Load (MW))

Conv Coal
7

HW CCGT
CCGT

4
100

150

200

250

300

350

400

450

Load [MWe]

Figure 8.6 First order derivative curves


With use of this first order derivative we get the differential heat per unit en with this differential heat
multiplied with the fuel cost the total differential- or marginal cost can be defined. Suppose the
conventional coal unit is already running full load and the HW CCGT and CCGT, both gas fired, are
running at a load with equal fuel cost. In case the load has to increase with 160 MWe the units have to
ramp up. For an economic optimum, it is a minimum of cost, the load has to be divided over the two
units in such a way that the marginal cost for each unit is equal and this is expressed with the
differential heat, because of the equal fuel cost, for the units in Figure 8.7.
7.0

1,750 GJfuel /h

1,460 GJfuel /h

6.5
d (Heat Consumption (GJ/h))/d (Load (MW))

850 GJfuel /h

6.0

y = 150 MW

1,400 GJfuel /h

Conv Coal

5.5

HW CCGT
CCGT
Setpoint

5.0

4.5
x = 10 MW

4.0
100

150

200

250

300

350

Load [MWe]

Figure 8.7 Differential or marginal fuel consumption curves for HW CCGT and CCGT

204

Chapter 8 Dutch Electricity Supply Model

As can be seen in Figure 8.7, the CCGT has to increase its load with 150 MW and the HW CCGT with
10 MW in order to equal differential heat consumption (CCGT (1750 - 850)/150, HW CCGT (1460 1400)/10). The differential heat curve for the CCGT is relatively flat and that is the reason why this unit
has to increase the generated load, with 150 MWe, 15 times the speed of the HW CCGT with 10 MWe.
In case the load has to decrease, this is opposite.
It is obvious that in case of considerably more than two units, of which many have to cover the
demand load curve, the difficulties with respect to an optimal economic choice highly increases. For
example, to commit a generating unit is to turn it on, that is to bring the unit up to speed,
synchronize it with the system, and connect it so it can supply power to the grid, this takes time. The
same counts in case a unit is switch off. From an economic point of view, losses occur at startup and
shutdown of a generating unit which also has to be taken into account. It is a problem of economics;
committing enough units and leaving them on is quite expensive. Lower total costs can be achieved by
turning off units (decommitting them) when they are not needed. Units that are still needed to meet the
system's demand can operate at a higher load and closer to their best efficiency. This is preferable
compared to having more units operating at a lower load and further from their best efficiency as this
would increase the net costs. However, it takes time to decommit a unit and a large unit is not switch
off for only one or a two hours but for at least 4 (CCGT) or even 24 hours (Conventional Coal); see
Section 4.7. The problem of economics as described can be solved by using Unit Commitment
function in combination with Economic dispatch models for Unit Commitment. Both functions are
described briefly in this section, an extensive description can be found in Wood et al (1996).
Unit Commitment
Unit commitment is an operations scheduling function which is sometimes called pre-dispatch. In the
overall hierarchy of generation resource management, the unit commitment function fits between the
economic dispatch and the maintenance and production scheduling. Unit commitment schedules the
on and off times of generating units, and finds the minimum cost hourly generation schedule while
ensuring that start up and shutdown rates, minimum up and minimum down times are considered.
Thermal System Modelling Considerations
The operational performance of thermal generating units in the system is represented in term of
generating cost, typically including fuel cost, maintenance cost and start up cost.
Fuel costs:
The fuel costs are represented by the total cost of generation from a facility as a function of the MW
power level. Fuel cost for generating at Pi MW for unit i can be represented by:

FCi ( Pi ) = Fi Hi ( Pi )
where:
FCi
Hi(Pi)
Fi

Fuel Cost for the unit (/h)


Heat consumption curve for the unit ((GJ/h)/MWh), see also Section 4.7
Fuel cost (/GJ)

Maintenance cost
The maintenance cost is typically represented in the form:

MCi ( Pi ) = BMi + IMi Pi


where:
MCi
BMi
IMi
Pi

Maintenance cost for the unit ()


Base cost ()
Incremental cost (/MWh)
Produced electricity (MWh)

205

Part 4

Start-up costs
The start-up costs vary between the maximum amount for a cold start to a much smaller amount for
a unit that was recently turned off and is still relatively close to its operating temperature. During its
down period, a thermal unit can be treated in two ways:
1. Cooling
2. Banking = hot stand by
In the case of cooling, the thermal unit is allowed to cool down after a stop and then heated back up to
the operating temperature over time for a scheduled turn on. In the case of banking, sufficient fuel
energy has to be put into the boiler regularly to maintain the operating pressure and temperature. The
start-up cost for a unit that is allowed to cool down first increases and then remains constant while the
unit is cooling down. The start-up cost for a unit that is banked is linear over time. After a few hours,
the start-up cost for the banked unit crosses the start-up cost for the cooled down unit (Figure 8.8).
The banking alternative is usually chosen when the unit is to be required again in a short period of
time.
Start-up
cost
Cooling

Banking

Cf

Figure 8.8 Time-dependent start-up costs after a stop


Start-up cost when cooling:

Ccooling = Cc (1

) FC + Cf

where:
Ccooling
Cc
FC
Cf
t
T

Start-up cost ()
Heat required for cold start-up (GJ)
Fuel costs (/GJ)
Fixed costs (includes crew costs, maintenance expensive ())
time the unit was cooled (h)
thermal time constant for the unit (h)

Start-up cost when banking:

Cbanking = Ct t FC + Cf
where:
Cbanking
Ct
t
FC
Cf

206

Start-up cost ()
Cost of maintaining unit at operating temperature (GJ/h)
Time the unit was banking (h)
Fuel costs (/GJ)
Fixed costs (includes crew costs, maintenance expensive ())

Chapter 8 Dutch Electricity Supply Model

There is no time-dependent component for a stand alone gas turbine because the work medium is air
instead of water and the specific heat constant of water is a factor 4 higher than for air. Also when a
gas turbine stops, surrounding air still flows through the compressor, combustor and turbine section
and equipment is cooling down faster than a boiler or HRSG.
A thermal unit can only undergo gradual temperature changes. This means that a few hours are
required to bring the unit on line. As a result of such restrictions, a number of unit constraints arise.
Unit Constraints
Thermal units are usually divided into the following types with their own characteristic constraints:

Must-run units:
Units that must be on-line for the whole year or certain periods of the year (if available), due to
operating constraints, reliability requirements, heat delivery obligations, fuel burning
obligations or economic considerations.
Cycling units:
Units that can cycle on and off and are subject to minimum up- and down-time constraints.
Both must-run and cycling units are dispatched economically between their minimum and
maximum load limits.
Peakers:
Units that can start up quickly and are typically not subject to minimum or down-time
constraints.

Summarized, the most important constraints which are applicable for the thermal units:
Maximum and minimum load
Units can only generate between given limits. Moreover, for selected hours, units may be unavailable
(due to planned maintenance or forced outages), or derated (reduced generating limits), or required to
operate at pre-specified generation levels for example due to heat or fuel burning obligations. The
operating limits of the generating units are represented by:

Pi min Pi (t ) Pi max
The minimum power of a unit may range between 10 to 25% of its rated capacity.
Minimum Uptime
The minimum uptime specifies the minimum time a unit must run before it can be shutdown. This is
largely a matter of judgment and are intended to provide time for temperature equalization within the
turbine and boiler to maintain stresses due to temperature differentials within safe limits.
Minimum Downtime
In the manner similar to minimum uptime, the minimum down time specifies how long a unit must be
down before it can be recommitted.
Crew Constraints
A plant that consists of two or more units cannot start-up more than one unit at a time because there
are not enough people to carry out all tasks
Spinning Reserve
The spinning reserve is the total amount of generation capacity available from all of the units that are
synchronized (i.e. spinning) on the system minus the current load and losses being supplied. The
spinning reserve must be carried out so that the loss of one or more units does not cause the system
frequency to drop too low.
Fossil-fuel usage
Fuel contracts may have limits and penalties for under-use and over-use of an allocated quota. In
case of gas as fuel also the maximum hourly capacity of gas can be limited.

207

Part 4

Within all these constraints and power demand to be served, the unit commitment function should
ideally find the feasible schedule for the available units which results in the least cost. The question
that is asked in the unit commitment problem is approximately as follows (Wood et al, 1996):
Given that there are a number of subsets of the complete set of N generating units that would satisfy
the expected power demand, which of these subsets should be used in order to provide the minimum
operating cost.
This unit commitment problem may be extended over some period of time such as the 24 h of a day or
the 168 h of a week. For a system of 100 units, and a time horizon of 168 hours one week, there are
already 30,000 decision variables which makes unit commitment a very large non linear mixed integer
programming problem. The real practical barrier in the optimized unit commitment function therefore is
the high dimensionality of the possible solution space. The most talked about techniques for the
solution of the unit commitment problem are (Wood et al, 1996):
Priority-list schemes;
Dynamic programming (DP);
Lagrange relation (LR).
The solution procedure involves the economic dispatch problem as a sub problem. That is, for each of
the units already connected to the system, the particular unit should be operated in optimum economic
fashion for the system. This will permit finding the minimum operating cost for that unit, but it does not
establish which of the units is in fact the one that will operating and will give minimum cost over a
period of time, this problem is handled within the unit commitment function.
Economic dispatch models for unit commitment
The optimum operation of all-thermal units can be classified as economic dispatch. The purpose of the
economic dispatch is to minimize the total operating costs of the units already connected to the
system while satisfying the power demand requirements of the system. Consider a system of N
thermal units connected to a grid with no restrictions serving a received electrical load Pload. The input
of each unit, shown as Fi, represents the units cost rate. The output of each unit, Pi, is the electrical
power generated by that particular unit. The systems total cost rate FT is the sum of the individual
units costs:

FT = F1 + F2 + F3 + ....... + FN
N

FT = Fi ( Pi ) = object function
i =1

= 0 = Pload Pi
i =1

The objective is to minimize FT subject to the constraint that the sum of the powers generated must be
equal to the power demand. The most effective way to solve the problem is to use the concept of
Lagrange multiplier and to minimize the augmented objective. This is known as the Lagrange
function L:

L = FT +
L dFi ( Pi )
=
=0
Pi
dPi
dFi

0=
dPi

208

Chapter 8 Dutch Electricity Supply Model

The incremental cost rates of all the units has to be equal to the undetermined Lagrange multiplier
and this objective is the necessary condition for the existence of a minimum cost operating condition
for the thermal units connected to the system. Two inequalities have to be added to this objective, the
constraint equation that the sum of the power outputs must be equal to the power demanded by the
load en second the power output of each units must be greater than or equal to the minimum power
permitted and must be less then or equal to the maximum power permitted on that particular unit.

dFi
=
dPi
Pi , min Pi Pi , max
N

P = P
i

load

N equations
2N inequalties
1 constraint

i =1

This set of nonlinear equations, which have to be solved, requires iterative equations for their solution
with use of for example:

The lambda-iteration method: This methods starts with an assumed value of lambda and finds
the corresponding power output of each generating unit from incremental cost characteristics.
The sum of generations is calculated and compared with the required power demand. In case
the lambda was chosen to low, the sum of generations will be less then the required power
demand and the lambda has to be increased for the next calculation. To solve the equitations
the lambda has to be adjusted up or down until the power balance equation is satisfied. With
results of two attempts, we can extrapolate (or interpolate) the two solutions to get closer to
the desired value of the power demand.

Gradient search method: This method works on the principle that the minimum of a function,
f(x), can be found by a series of steps that always goes in a downward direction. To solve the
economic dispatch problem which involves minimizing the objective function and keeping the
equality constraint, the gradient technique must be applied to the Lagrange function it self.
Other methods which can be used are:
Newtons Method;
Economic dispatch with piecewise linear cost functions;
Economic dispatch using Dynamic Programming;
Base Point and participation factors;
Custom made programs.
For a more extensive description of the lambda-iteration-, gradient- and the other methods but also for
the solution of the unit commitment problem mentioned previous please refer to Wood et al (1996) and
Delarue & Dhaeseleer (2007).
PowrSym3TM as Solver of Unit Commitment / Dispatch problem in model
As stated at the begin of this Section we have to find a method to calculate the optimum economic
load division of the grid coupled individual generation units, with the applicable constraints, for the
electricity generation companies' units in our model. We used the PowrSym3TM computer model
conceived by Operation Simulation Associates (OSA) for that. PowrSym3 is a multi-area, multi-fuel,
chronological production cost simulation model for electric power systems and combined heat and
power systems. This model enables us unit commit/dispatch of the modeled conventional- and
combined heat/power units in our system in such a way that the expected power and heat demand is
matched against a minimum of operating cost (optimum economics). Although we tried first to create
our own dispatch model in Matlab, this was stopped due to the limitations it brought with it in particular
in the area of committing heat/power units. Together with and thanks to OSA and Mr. J.Th. Kromhout
from E.ON Benelux, a derived PowrSym3 model was built for the unit commitment/dispatch for this
research project. In Figure 8.9 the data flow diagram of the PowrSym3 module (as part of the main
model in Figure 8.2) for this study is given. The input database file is prepared directly in MS Excel
and imported in PowrSym3. The PowrSym3 output data file is exported to a MS Excel sheet.

209

Part 4

Pow er
System Data

Database
MS EXCEL

Hourly Load
Data

Weekly
Maintenance
Data

Powr Sym3

Reports

Hour

Week

Month

Calender
Year

Figure 8.9 Data flow diagram of the PowrSym3 computer model for this research project
PowrSym3TM is a commercial available package of which an extensive description as well as the major
features are given in annex C.
For the purpose of this thesis not all the mentioned features of PowrSym3TM as described in the annex
where required. For example, transmission constraints are not modeled, the fuel supply is unlimited
(no constraints), one control area is simulated and 8,760 hour simulation is used starting on a Monday
(internal calendar of PowrSym3). Of course the other features can be used in case of derived studies
where these other features are required.

210

Chapter 8 Dutch Electricity Supply Model

8.4.2

Economics of Other Electricity Producers, Combined Heat/Power, Waste Incinerators, Nonfossil

Combined Heat/Power
The amount of generated electricity, produced heat, required fuel input and CO2 emissions of
combined heat and power is calculated in the Industrial CHP, DH CHP and Small CHP sub-modules
as blocks, on hourly basis; the results are used as input for the main model. The applied commitment
methodology in the CHP sub-modules is explained below.
The physics and economics of (existing) heat and power units are determined by a combination of
income and costs (see Figure 8.10).
Electricity Grid

Supply to Grid

Gas Grid

Operation Buyer
Electricity
CHP - Plants + Boiler(s)

E-purchase
Heat
Own Consumption

Buyer

Figure 8.10: Physical streams and economics of combined heat and power plant
Taking a closer look at the profit and loss account of an average combined heat and power plant, we
see the following distribution of income and expenditures:
9

Income
- Electricity income in peak, off-peak (including subsidies)
- Heat sales (process steam and/or DH heat)
- Avoidable cost2 compensation (applies primarily to industrial buyers)

Expenditure
- Purchase of natural gas (commodity, energy tax, capacity costs LNB, RNB)
- Maintenance costs (variable and fixed)
- Personnel costs
- Reserve position electricity-grid (grid operation), electricity purchase costs
- CO2 rights
- Depreciation

Operating result = Income - Expenditure

Result = Operating result - Interest paid

Avoidable cost are an income based on previous variable and fixed cost of the older CHP installation on the
CHP site. Mostly these older units were sized on the heat demand of the location only. This income was
applicable for CHP units placed in Joint Ventures between the industry and energy companies.

211

Part 4

For the commitment of a combined heat and power unit, the variable costs in combination with
electricity revenues (including subsidies) are important. The fixed costs have no impact, although they
do contribute to an important degree to the result. The important variable costs are:
9
9

Natural gas commodity costs including energy tax (if applicable)


Variable maintenance costs

On average, the natural gas commodity costs represent approximately 60 to 70% of the total costs
(variable + fixed). The cost of required CO2 credits (applicable as of 2004) for the medium and small
sized CHPs are assumed to be more or less fixed in a year caused by the large amount of operating
hours and play no role in the decision for part or full load in the model calculations.
To explain the methodology of commitment of a combined heat and power plant in peak- and off peak
periods the methodology is first described and in the next part worked out in more detail for the
following CHP production options:

Combined Cycle Gas Turbine (CCGT), 2 cases.


Gas turbine (GT) + Heat Recovery Steam Generator (HRSG), 2 cases.
Boiler + Steam Turbine (ST), 1 case.
Gas engines, 1 case.

Description methodology of commitment for a combined heat and power plant


The CCGT and GT + HRSG units can be powered down during nightly off-peak periods and on
weekends in case the variable costs exceed the income. Units are rarely shutdown completely during
the off-peak period because other considerations such as the security of supply of heat play an
important role. Frequently starting and stopping units affects the availability and reliability of a unit and
can lead to unexpected failure as a result of additional wear and tear on the parts. It also affects the
residual life expectancy of the unit. The failure to supply heat has large financial consequences for the
industry to which the steam is supplied (paper processing plants, chemical plants, etc.). The 8 hours of
the nightly off-peak period are too short to shut an installation down, start up one or more auxiliary
boilers, and start up the unit again later and shut down the auxiliary boilers. The weekend has a
slightly longer off-peak period (56 hours), which is why it can make more sense to shut down a unit
then. Again, also during a weekend stop, the availability and reliability of a unit is affected but also,
because of the longer off-peak period, the financial aspects must be considered.
As mentioned in Section 8.3.5, CHP plants of the other electricity producers sell their electricity often
to the trade department of electricity production companies. The natural gas can also be bought at
these trade department at the same time. The development of the combination of electricity- and gas
forward prices, also called spark spread (explained hereafter), is of importance to judge if the margin
between electricity income and gas cost is acceptable.
To determine the fuel cost of electricity generation for a CHP unit, the costs of heat can be (based on
a 90% efficiency of a typical boiler for separate heat production) deducted from the total fuel costs for
the unit. After deducting the fuel for heat of the total fuel input, the fuel required for electricity
generation remains. The electricity generated divided by this fuel gives the arithmetical or spark
spread efficiency. Mathematically expressed:

eArith =

e
th
1
th boiler

where:
eArith
e
th
th boiler
212

The arithmetical efficiency or spark spread efficiency


The electrical efficiency of the CHP
The thermal efficiency of the CHP
The thermal efficiency of the (reference) boiler for separate heat production

Chapter 8 Dutch Electricity Supply Model

The spark spread efficiency depends on the unit type, units electrical load and units heat load. For a
larger CHP unit operating in full load, this means spark spread efficiencies between 54% and 76% are
possible, 54% with no heat load and 76% with maximum heat load. In practice for the different CHPs
types, the cost of heat (and income for the seller) is a wide range and depending on the type of
contract between seller and buyer of heat.
With the spark spread efficiency and the fuel cost, the fuel cost of electricity generation of a CHP unit
can be calculated. In combination with the electricity prices for base, peak or off peak, the spark
spread (difference between electricity price and the fuel cost) for base, peak and off peak can be
calculated. The spark spread development, based on the forward electricity prices (for peak and off
peak) and gas prices for 2003 (see Figure 2.2 of Chapter 2) is given in Figure 8.11 for a spark spread
efficiency of 54% and 76%.
35

30

25

Sparkspread [/MWh]

20

15

10

27-12-02

12-12-02

27-11-02

12-11-02

28-10-02

13-10-02

28-09-02

13-09-02

29-08-02

14-08-02

30-07-02

15-07-02

30-06-02

15-06-02

31-05-02

1-05-02

16-05-02

1-04-02

16-04-02

17-03-02

2-03-02

15-02-02

31-01-02

1-01-02

-5

16-01-02

Peak 76%
Peak 54%

-10

Off Peak 76%


Off Peak 54%

-15

Figure 8.11 : Calculated forward spark spread prices in 2002 for 2003
As can be seen in Figure 8.11, the forward spark spread for off peak for 54% spark spread efficiency
(no heat load) is negative i.e. every MWh generated electricity gives a negative financial result. For the
76% spark spread efficiency (maximum heat load) the forward spark spread is approximately zero i.e.
every MWh generated electricity gives nor positive nor negative financial result. The forward spark
spread for peak however is positive for both 54% and 76% spark spread efficiency, but clearly with
76% spark spread efficiency more financial result is achieved.
The use of the spark spread is nowadays common for contracting electricity by electricity producers
because it decreases the risks of an increase or decrease of gas prices during the year of operation
whilst the electricity price is already fixed the year before. In a contract based on a spark spread the
moment the electricity prices are fixed, this also counts for the commodity gas prices, the margin is
fixed. During the year of operation, the margin (= spark spread) between electricity price and fuel cost
stays the same; in other words, if the gas price increases also the electricity prices increases and vice
versa.
With the combination of forward spark spread development for peak and off-peak and the knowledge
of the limits of range in which the CHP unit can operate, the producer can calculate the profit
(electricity income) and loss (gas cost) at each moment and choose the most optimum momentum of

213

Part 4

contracting. With a positive spark spread for peak- and negative spark spread for off peak hours in
2003 this means for CHP based on CCGT and GT + HRSG full load operation during peak and part
load or even shutdown of the unit in off peak periods.
In the model calculations, the spark spread for off peak periods depends on the fuel prices for coal
and natural gas based on the applicable scenario. In case coal prices are lower than gas prices and
conventional coal-fired units can generate most of the off peak electricity (depending on the installed
conventional coal capacity), the spark spread for off peak is negative for CHP based on CCGT and GT
+ HRSG units. In this case these units generate a minimum of the off peak electricity depending on the
part load level which can be reached at a given heat demand but units are not shut down because of
security of supply reasons for heat as mentioned before. In case the coal prices are slightly less, equal
or even higher than gas prices, the spark spread for off peak is positive for CHP based on CCGT and
GT + HRSG units and these units generate most of the off peak electricity and conventional coal-fired
units generate a minimum of the off peak electricity depending on the part load level which can be
reached.
Boiler and steam turbine (ST) units follow the heat demand, and electricity generation is closely linked
to the heat demand, part load operation, due to negative spark spreads are therefore not an option.
Gas engines are usually flexible because the heat is often supplied as hot water for indoor heating of
greenhouses, residences and/or offices. The nature of this sort of heat supply, especially in
combination with heat buffers and/or backup boilers, makes gas engines excellent start/stop
machines. They are usually run at full capacity during the peak periods and are shutdown during off
peak periods in case variable cost exceed the income (negative spark spread) and heat buffers are
available.
Methodology of commitment for a combined heat and power plant worked out for six cases
The unit type and the units limits of range for the electrical load and heat load are of importance for
the commitment. To show the differences, several cases are worked out for the reference situation in
2003. The following assumptions for the reference situation, used technology, applied electricity- and
gas prices and other issues were made:
Reference situation for CHP:

As reference unit for electricity, a modern CCGT based on the GE F9A with an efficiency of 54%.

As reference for the supply of steam, a steam boiler with a fixed assumed efficiency of 90%.

As reference for hot water, a hot water boiler for heat and hot tap water with a fixed assumed
efficiency including of 90%.
Technology:

Heat consumption curves and energy balances at part load are determined using the GT
Pro/GTMaster program.

The energy balance (fuel, electricity and heat) for all possible operational (full- and part load)
situations.

Design data is used for the efficiencies calculations.

Maintenance cost are assumed to be fixed, caused by the large scale of possible contract options
for maintenance in the market nowadays.
Applied prices:

Average forward prices in 2002 for base-, peak- and off peak electricity prices 2003 (Base 31.4
/MWh, Peak 47.6/MWh and Off peak 17.4/MWh)

Average forward prices in 2002 for natural gas commodity prices for 2003 ( ct 12.0/m3)

Development of average spark spread for 2003 based on above forward prices and selection.

MEP subsidy of 5.7/MWh (until 1000 GWh generation), during 2003 replaced by CO2 index for
CHP for 2003.

Heat revenues based on the reference situation with the forward average natural gas commodity
prices.

Forward prices CO2 for 2003 are not applicable.

214

Chapter 8 Dutch Electricity Supply Model

Description of the Cases:


The below cases are elaborated using the assumptions described above (see Annex D for extensive
details of the calculations):
1.
2.
3.
4.
5.
6.

A modern large-scale CCGT with 1 GE 9FA gas turbine, HRSG without supplementary firing and
a steam turbine with the possibility of 100% condensation operation.
A smaller CCGT with 2 ABB GT10 gas turbines, HRSG with supplementary firing and a steam
turbine.
A 5 MW Typhoon gas turbine with supplementary fired HRSG and no steam turbine.
A GE LM 6000 MW with supplementary fired HRSG and no steam turbine.
A conventional steam boiler with back-pressure turbine.
A gas engine of approximately 500 kW with hot water supply only.

Outcome of the Cases:


For the six mentioned cases the following outcomes are calculated for several electrical and heat load
situations:

The fuel cost of electricity generation per MWh (corrected for proceeds from heat);

The energy savings in % with respect to the fuel consumed for separate generation with the
reference CCGT and the fuel consumed for separate heat generation with the reference boiler;

The same as above for CO2 emission reduction;


In the summary of this sub-section for CHP the cost prices of generation for the specified load
situations of the cases are compared with forward electricity prices for off peak and peak hours. The
differences between the cases are discussed and influences of MEP subsidies explained.
Option 1:
The commitment range for a modern large-scale CCGT with 1 GE 9FA gas turbine without a
supplementary fired HRSG and with 100% condensation operation is displayed in Figure 8.12.
90

400

80

70

60

Power [MWe]

300

50
250
Limit of operating range

40

30

200

20
150

Electrical, Thermal and Total Efficiency [%]

350

GT 100%
10

GT 80%
GT 50%

100
0

50

100

150
Heat Load [MWth]

200

0
250

E - EFF (GT 100%)


Th - EFF (GT 100%)
Tot - EFF (GT 100%)

Figure 8.12: CCGT with 1 GE 9FA gas turbine, HRSG and steam turbine with 100% condensation
operation
The electricity loss that results from supplying heat is about the same for all of the gas turbine loads.
For this reason, the loss factor for such large-scale plants may be assumed to be constant.
Figure 8.13 displays the savings in percent compared with the reference situation, and the fuel cost of
electricity generation.

215

20

50

15

45

10

40

35

30

-5

25

-10

20

-15

15

-20

10

-25

-30

Fuel Cost E (corr. for W) [/MWh]

Saving Fuel vs Reference[%]

Part 4

Savings GT100%
Savings GT 80%
Savings GT 50%

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Cost E GT100%
Cost E GT80%

1.0

Cost E GT50%

Heat / Power Ratio [-]

Figure 8.13: Savings in percent with regard to the reference; fuel costs per MWh corrected for heat
The figure clearly demonstrates that the fuel savings increase when the heat/power ratio increases
and the GT is running at full load instead of part load. The fuel cost of electricity generation increase
when the plant is operated at part load. The choice to run the plant at part load is therefore made
based on absolute income and costs in relation with the required heat.

90

90

80

80

70

70

60

60

50

50

40

40

30

30

20

20

10

10

Electrical, Thermal and Total Efficiency [%]

Power [MWe]

Option 2:
The commitment range for a smaller CCGT with 2 ABB GT10 gas turbines, HRSG with supplementary
firing and a steam turbine is displayed in Figure 8.14.

GT 100% + SF
GT 100%
GT 80%
GT 50%

0
0

10

20

30

40

Heat Load [MWth]

50

60

70

E - EFF (GT 100%)


Th - EFF (GT 100%)
Tot - EFF (GT 100%)

Figure 8.14: Small-scale STEG with 2 ABB GT10 gas turbines, HRSG with supplementary firing and a
steam turbine

216

Chapter 8 Dutch Electricity Supply Model

20

50

15

45

10

40

35

30

-5

25

-10

20

-15

15

-20

10

-25

-30

Fuel Cost E (corr. for W) [/MWh]

Saving Fuel vs Reference[%]

The electricity loss that results from supplying heat is about the same for all of the gas turbine loads
up to the point at which supplementary firing gets involved. For this reason, the loss factor for such
smaller plants may be assumed to be constant as long as the required heat can be generated without
supplementary firing. Figure 8.15 displays the savings in percent and the fuel cost of electricity
generation.

Savings GT 100% + SF
Savings GT 100%
Savings GT 80%
Savings GT 50%
Cost E GT 100% + SF
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

Cost E GT 100%

1.3

Cost E GT 80%

Heat / Power Ratio [-]

Cost E GT 50%

Figure 8.15: Savings in percent with regard to the reference; fuel costs per MWh corrected for heat
Option 3:
The commitment range for a 5 MW Typhoon gas turbine with supplementary fired HRSG and no
steam turbine is displayed in Figure 8.16:

35

100

90

80

Power [MWe], Fuel [MWth]

25

70

60

20

50
15

40

30

10

20
5

Electrical, Thermal and Total Efficiency [%]

30

GT 100%
GT Fuel

10

0
5

10

15
Heat Load [MWth]

20

25

Total Fuel
E - EFF (GT 100%)
Th EFF (GT 100%)
ToT EFF (GT100%)

Figure 8.16: Typhoon 5 MW gas turbine with supplementary fired HRSG and no steam turbine

217

Part 4

20

50

15

45

10

40

35

30

-5

25

-10

20

-15

15

-20

10

-25

-30

Fuel Cost E (corr. for W) [/MWh]

Saving Fuel vs Reference[%]

For these plants, the electricity loss is not the result of supplying heat because only the gas turbine is
supplying electrical power. If more heat is required, the supplementary firing of the HRSG is turned up.
The increase in fuel consumption (see Figure 8.17) can be fully attributed to the heat. Such smallscale GT plants are not usually powered down, which is why only the GT 100% case has been worked
out.

0
1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Savings GT100%
Cost E GT 100%

Heat / Power Ratio [-]

Figure 8.17: Savings in percent with regard to the reference; fuel costs per MWh corrected for heat
Because the supplementary firing efficiency is higher than the reference efficiency for heat, the
electricity generation costs decrease when the heat/power ration increases, see Figure 8.17.
Option 4:
The commitment range for a GE LM 6000 gas turbine with supplementary fired HRSG and no steam
turbine is displayed in Figure 8.18:
90

140

80

70

Power [MWe], Fuel [MWth]

100
60

80

50

40

60

30
40

Electrical, Thermal and Total Efficiency [%]

120

GT 100%
20

GT 80%
GT 50%

20
10

GT 100% Total Fuel


GT 80% Total Fuel
GT 50% Total Fuel

0
20

25

30

35

40

Heat Load [MWth]

45

50

55

E - EFF (GT 100%)


Th - EFF (GT 100%)
ToT - EFF (GT 100%)

Figure 8.18: GE LM 6000 gas turbine with supplementary fired HRSG and no steam turbine

218

Chapter 8 Dutch Electricity Supply Model

As in option 3, the Typhoon gas turbine, the electricity loss is not a result of supplying heat as only the
gas turbine is supplying electrical power. The LM 6000 gas turbine is an excellent machine that is
easy to control (see Figure 8.18). Here too, the increase in fuel consumption can be fully attributed to
an increase in heat demand. The steepness of the supplementary fired curve is about the same for
each GT load, which implies that the HRSG does not lose efficiency when the gas turbine supplies
less heat.
20

40

15

35

10

Saving Fuel vs Reference[%]

25

-5

20

-10

15

-15

Fuel Cost E (corr. for W) [/MWh]

30
5

10
-20
Savings GT100%

-25

Savings GT 80%
Savings GT 50%

-30

Cost E GT100%

0
0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

2.2

Cost E GT80%

Heat / Power Ratio [-]

Cost E GT50%

Figure 8.19: Savings in percent with regard to the reference; fuel costs per MWhe corrected for heat
Because the supplementary fired efficiency is slightly lower than the reference efficiency for heat, the
fuel cost of electricity generation increase (see Figure 8.19) when the demand for heat increases.

200

100

180

90

160

80

140

70

120

60

100

50

80

40

60

30

40

20

20

10

Electrical, Thermal and Total Efficiency [%]

Power [MWe], Fuel [MWth]

Option 5:
The commitment range for a conventional steam boiler with back-pressure turbine is displayed in
Figure 8.20:

BPST
Boiler Fuel
E - EFF

0
0

20

40

60

80

100

120

Heat Load [MWth]

0
140

Th - EFF
Tot - EFF

Figure 8.20: Conventional steam boiler with back-pressure turbine


219

Part 4

20

50

15

45

10

40

35

30

-5

25

-10

20

-15

15

-20

10

-25

-30

0
4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

Fuel Cost E (corr. for W) [/MWh]

Saving Fuel vs Reference[%]

Because the conventional steam boiler and back-pressure turbine follow the heat demand and the
generation of electricity follows the heat demand, part load operation is not an option. In this case, the
additional boiler fuel consumed cannot be fully attributed to heat because additional electricity is also
generated.

Savings
Cost E

Heat / Power Ratio [-]

Figure 8.21: Savings in percent with regard to the reference; fuel costs per MWhe corrected for heat
Option 6:
There is no specific commitment range for gas engines as they are either switched on or switched off
(see Section 4.5.5). Table 8.3 shows the most important characteristics of a 0.5 MWe gas engine with
heat supply at two temperature levels. Without heat supply, the fuel cost of electricity generation are a
function of the electrical efficiency.

90C / 110C
70C / 90C

Net Power
[MWe]

Net Heat
[MWth]

Savings
[%]

Cost E
[/MWh]

0.5
0.5

0.7
0.8

17
21

32.0
28.0

Table 8.3: Gas engine characteristics


Summary
For the CCGTs CHP it is obvious that a larger plant (option 1) has less electricity production fuel cost
then a smaller plant (option 2). In both options however the cost of electricity does not come at all
above the average level of the off peak prices in 2003 (with fuel cost for heat based on 90% thermal
efficiency) but the differences between off peak prices and electricity production cost for the larger
plant are smaller. The absolute losses of revenues per hour during off peak for both options are lowest
in case of near maximum heat delivery and 80% part load of the gas turbine of the larger plant and
50% of the smaller plant. In case the MEP subsidy is included this leads to a shift towards more gas
turbine load.
The GT + HRSG CHPs can be divided in a larger (option 4) and smaller (option 3) the same as the
CCGTs CHP, but with the difference that the fuel cost of electricity generation for the large GT+HRSG
does not come at all under the level of the off peak prices in 2003 whilst for the smaller GT+HRSG

220

Chapter 8 Dutch Electricity Supply Model

with high heat/power ratio this is not a problem and fuel cost of electricity generation are lower than off
peak prices. In case the MEP subsidy is included this leads to a shift towards more gas turbine load.
The Boilers+ST option with high heat/power ratio is heat load following and fuel cost of electricity
generation are constant and lower than off peak prices.
The gas engines have fuel cost of electricity generation lower than the off peak price as well, but
operating this engines is more or less a function of the heat demand and this is fluctuating over the
seasons.
As described before, in the model calculations the spark spread for off peak periods is depending on
the fuel prices for coal and natural gas based on the applicable scenario. With this spark spread also
the amount of generated electricity during off peak of the CCGTs, GT+HRSG and gas engines CHPs
is calculated. The supply of the larger CCGTs CHP with an installed capacity above 250 MWe is
calculated in the PowrSym3 module in our model. It is assumed that these large units are part of the
unit commitment/dispatch calculation, with the applicable heat delivery obligations, of the electricity
production companies.
Non-fossil
In this model, the non-fossil sub module is divided into an arithmetical model and a block approach. A
arithmetical model approach is used for on- and offshore wind power. A block approach is used to
calculate the generated electricity with hydropower, photovoltaic solar energy and small scale biomass
incineration and biomass fermentation. In the block approach, installed capacity is multiplied with a
capacity factor resulting in a supplied power.
Wind power
The arithmetical approach is used for wind power generation, whereby a distinction is made between
wind power that is generated onshore and offshore.
Onshore
The wind-power production model of Ter Borg (2005) is used to calculate the onshore wind power
generation. The wind data (wind speed (see Figure 8.22) and wind direction per hour) for 2001 was
downloaded from the KNMI database for four weather stations Schiphol, De Kooy, Stavoren, and
Valkenburg and fed into the wind-power production model. The wind power generation was
determined for an installed wind power capacity of 1,000 MWe, which is evenly distributed across the
locations near the previously mentioned weather stations. The installed wind power capacity is scaled
up or down depending on the applicable scenario. The correlation between the wind speeds measured
at each station and the recorded wind energy generation (Ter Borg, 2005) allows this. The capacity
factor (real wind generation (MWh)/full load (8,760h) wind generation (MWh)) results in approximately
0.25.

221

Part 4

1000

25.0

900

20.0

700

600

15.0

500

400

10.0

Windspeed @ 10m [m/s]

Wind Energie Production [MWhe]

800

P(v)
Stavoren
De Kooy
Schiphol
Valkenburg
Noordwijk

300

200

5.0

100

0
1-jan-01

0.0
1-feb-01

Figure 8.22: Wind speed and wind energy generation at four weather stations for 1,000 MWe wind
power for the month January
Offshore
The calculation method of Boer (2004) was used to create our own model to calculate the offshore
wind power generation. The model was first created for the offshore wind farm at Egmond aan Zee
where 36 Vestas V90 windmills with a total installed capacity of 108 MWe will be built approximately
10 - 18 km off the coast. The model can easily be extended for more installed capacity providing the
capacity curve for the selected wind turbines is known. The 2001 wind data for Noordwijk was used to
calculate the wind energy generation. The wind speed is measured at a height of 27.6 m and
corrected to a wind speed measured at a height of 10 m above open water with a roughness length of
0.002 m also called the potential wind (KNMI).
The Vestas V90 turbines have a hub height of approximately 65 m. This means that the potential wind
speed at 10 m needs to be corrected to a wind speed at this height. This is done using the MoninObukhov theory, which describes the boundary layer of the wind velocity profile between the blending
height (approx. 60 m) and the surface (0 m). At heights above 60 m (the blending height) the wind
speed virtually depends on the height and hence remains constant. The layer above the blending
height is called the Ekman layer. It is assumed (Boer, 2004) that the Monin-Obukhov theory can be
used to calculate the wind speed because the rotor height is just above the blending height.
The equation for the wind speed at the blending height (approx. 60 m) and at the measured height (10
m) is as follows:
v_blend = v_measured * ln ( z_blend/ z_0) / ln ( z_measured/ z_0),
with z_0 = 0.002 m
This results in a correction factor of the wind speed at 10 m of approximately 1.23.
The 2001 wind data for Noordwijk was used to calculate the wind speed for one year for the previously
mentioned 108 MW wind farm equipped with 3 MWe Vestas V90 turbines. Without availability or other
corrections, the capacity factor is 0.466. It will, however, be lower once the corrections for farm

222

Chapter 8 Dutch Electricity Supply Model

availability, farm efficiency (the negative effect the turbines have on each other), and the grid losses
have been applied (see also ECN, 2002). For a farm availability of 98.5%, a farm efficiency of 80%,
and a grid loss of 5%, the capacity factor is 0.35. The wind speed and wind energy generation curve
for a week is displayed in Figure 8.23.
100

25

90

80

20

60

15

50

40

10

Windspeed @ 65 m [m/s]

Wind Energy Production [MWhe]

70

P(v)
V blend

30

20

10

0
1-feb-01

1-jan-01

Figure 8.23: Wind speed and wind energy generation at Noordwijk weather station for 108 MWe wind
power for the month of January
Hydropower
For hydropower, the variations in river flow amounts could be taken into account, but because of the
present and expected future small scale of hydropower in the Dutch situation an average load is taken
into the sub modules.
Photovoltaic solar energy
For photovoltaic solar energy, sunshine is taken into account for a standard year. This creates an
annual distribution as displayed in Figure 8.24.
400

350

Solar radiation [Wh/m 2]

300

250

200

150

100

50

8500

8000

7500

7000

6500

6000

5500

5000

4500

4000

3500

3000

2500

2000

1500

1000

500

Figure 8.24: Solar radiation for a standard year

223

Part 4

Small scale biomass incineration and biomass fermentation


The installed capacity for this type of generation is like hydropower very small and therefore an
average load is taken into the sub modules. The co-fired biomass combustion in the large production
units is included by the electricity generation companies.
Waste Incinerators
A waste incinerator's primary task is to process waste in an environmentally responsible manner. Most
of the large incinerators reclaim the heat that is released and use it to generate electricity. Some of the
incinerators also supply process and/or district heat. The economics of waste incinerators are mainly
determined by the quantity of waste incinerated. A rate, which takes into account the incinerator's
income and costs, is paid for every ton of waste delivered to a waste incinerator. In other words, the
rate is a balancing item. The additional (derived) income from electricity, heat (if applicable), MEP (if
applicable) and residues (see Figure 8.25) impacts the height of the rate paid per ton of waste
delivered. Higher electricity rates or higher income from residues, for example, normally result in a
lower rate per ton of delivered waste.
The Dutch Waste Management Association (Vereniging van Afvalbedrijven (VAV)) assumes a (gross)
energy efficiency from waste, which was an average of 520 kWh per ton of incinerated waste in 1996
(VAV, 2004). More recently, the energy efficiency amounted to an average of 630 kWh/ton waste. In
this thesis, a conversion factor was applied to the high-pressure steam supplied by AZN in Moerdijk to
include it in the amount of electricity generated.

Electricity Grid

Supply to Grid

Waste Market

Buyer Operation
Waste Incinerator
(Waste, , Availability, MEP)

Electricity
E-purchase
Heat
Own Consumption

Supply to Heat Grid


Buyer
Heat Grid

Figure 8.25: Physic streams and economics of a Waste Incinerator


The model uses as starting points for the reference an installed electrical capacity of 414 MWe for the
scenarios. Expanding the incineration capacity also increases the installed electrical capacity.
Together with the availability of the incineration lines and supply in peak/off-peak periods (differences
are assumed to be small and especially based on the premise that maintenance is mainly carried out
during off-peak hours), this results in a gross energy efficiency that can be compared to the VAV data.
The own consumption percentage is used to determine the net electricity generation.
8.4.3

Economics of Import and Export

The economics of import and export are determined by a number of factors:

224

Chapter 8 Dutch Electricity Supply Model

The available interconnector capacity.


The difference in base and peak load prices with surrounding countries.
The cost of the interconnector capacity.
In case of import the availability of production capacity for export in surrounding countries and in
case of export the availability of inland production capacity.

These factors are described in detail in Van Eck et al (2004). The most important conclusions are as
follows:
The difference between base and peak load prices with the surrounding countries is the driving
force behind import and export.
A level playing field is absent caused by differences in deregulations of electricity markets, rates
for grid access, government incentives in relation to fuel policy, encouragement for renewable
energy and heat and power etc.
The Netherlands has a high percentage of gas-powered capacity of which fuel cost are high
(situation 2003) compared with coal and nuclear powered capacity in the surrounding countries.
The Netherlands is importing around 20% (situation 2003) of the required electricity from Germany
and France (see also Figure 9.21 of Chapter 9).
The advantages of lower base and peak load prices compared to the surrounding countries are
canceled out by the auction-related costs for interconnector capacity.
The Netherlands geographic location with respect to available cooling water of the North Sea and
the extensive high voltage grid with interconnectors (see Chapter 5) gives it an excellent starting
point as export country providing there is a level playing field in the near future.
The model uses a block approach for import and export loads. The load size and price of import can
be included in the dispatch as a virtual power plant. The composition and emissions of this virtual
power plant is based on the share of lignite, coal, nuclear and other generation capacity of the
exporting country. The load size for export can be added to the supply curve (see Section 8.3.4) and is
generated by the available inland generation capacity. The amount of imported or exported electricity
depends on the selected scenario (see Chapter 7).
8.5

Validating the Model

In order to validate the model, the reference year 2003 is simulated to see what differences in results
the model gives compared with the real situation in 2003. The overview of the results for 2003 are
given in annex F.
The load curve for electricity is based on the actual load curve of 2005 published by TenneT. With use
of a load conversion program this curve is converted to the 2001 calendar because of the internal
calendar of PowrSym3, see also Section 8.2. The load curve does not present all the generation and
therefore the missing decentralized generation is added in order to match the total electricity supply for
2003. As will be explained in Chapter 9, the choice for 2005 is because of the availability of more
TenneT data compared with 2003.
The load curves for the district heating are based on the load shape curves of 2005 of three larger
district heating areas (Amsterdam, Utrecht and Purmerend). With use of a load conversion program
this curves are also converted to the 2001 calendar. With the climate data of 2003 the curves are
sized in order to fit the production heat output of 2003.
With use of the sub modules the electricity and heat supply of the other electricity producers is
simulated for 2003. With the applicable prices in 2003 for electricity, coal and natural gas and the real
produced process/district heat in 2003, the technical input parameters (availability factor, peak/off
peak load %) of the sub modules are adjusted in order to fit the outcomes with the real situation in
2003. For wind supply (only onshore in 2003), the wind conditions (speed) of 2001 of 4 on land
stations is used in combination with the average installed wind power in 2003 to approach the same
wind supply as the real supply. The real 2003 wind data was not available as input for the wind-power
production model the moment it was used for the calculations.

225

Part 4

The outcome of the generated load in the sub modules and the real imported electricity load of 2003
published by TenneT, is subtracted from the total load curve and the resulting curve is the input supply
curve for the electricity production companies. With this supply curve, the curves for process and
district heating and the applicable constraints, the electricity and heat supply of the individual existing
power plant units modeled in PowrSym3 are simulated.
The last simulation step for the electricity production companies gives differences compared with the
real situation in 2003 (see Table 3.1 (Electricity), Appendix B (Fuel) and Section 2.2.1 (CO2)). The
main conclusions are as follow (model versus real) : for the produced electricity this is a deviation of
+1.0%, for the required fuel a deviation of -1.7% and for the CO2 emissions a deviation of +1.6%. The
deviations are relatively small but they can be explained as follows:

The PowrSym3 module calculates an optimum on the basis of minimum cost for the complete
supply of electricity generating companies, in real situation this optimum is determined by
every individual supplier and this leads to sub optimization (see also Van Eck et al, 2002)
compared with the results from our model.
The PowrSym3 module works with forecast percentages for planned maintenance and forced
outages. For this thesis, these forecast percentages are chosen equal for all individual units
because no real data was available while in the real situation these percentages are different
for every unit. For example, in 2003 the unit Amer 9, was shutdown for several months
because of an accident the moment the unit was in maintenance. The real percentage for
forced outage for this unit in 2003 was much higher compared with applied percentage in the
PowrSym3 module. In the PowrSym3 outcomes therefore much more conventional coal
electricity is generated including more CO2 emissions.
The simulated total electricity load is not the real electricity load; complete data for this load
are just not available. Measurement data of this load supplied by the grid operators and
published by TenneT was incomplete for 2003 and partly incomplete for 2004. After 2004 the
published load data became much more complete and represents almost 85 - 90 % of the
total supply (see also Section 9.2.3). In the model, the missing load is added to the measured
load curve to create a total electricity load curve.
The total net generated load of the electricity generation companies in the model is matching
with the CBS data (Section 8.3.3). In the real situation, the production data of each unit given
in the environmental reports of the units, can be used and gives insight in the type of supply in
the total net generation. These data, summarized to total net generated load of the electricity
generation companies do not match completely with the CBS data.
The APX day ahead market is used by electricity generation companies to sell and buy
electricity in case they are short or long the day ahead of the real delivery. Other electricity
producers also use the APX for selling extra available quantities however their share is less.
This day ahead market influences has also its effects on what (a number of) production units
really produces compared with the model results which does not work with day ahead
optimizations.
The imbalance market (intraday) can be used by electricity generation companies to settle
imbalances during the day of delivery. Besides the imbalance market every electricity
generation company can use its own optimization of its own portfolio. The model does not
work with imbalance and does not look at own optimizations but treats the system as a total.

In the scenario analysis, the calculated model results for 2003 are used as a reference. In this way all
cases, reference and future scenarios, have the same calculation basis.
8.6

Summary

A long term technical - economical simulation model for the energy demand and supply system was
developed. In this model the existing energy supply systems can be combined with possible future
energy supply systems based on the chosen scenarios. The outcomes of the model calculations for
the defined scenarios can be used to analyze the balance between the environment, economics and
security of supply.

226

References

References
Algemene Energie Raad AER (2004), Behoedzaam stroom opwaarts, ISBN 90 74357 385 [Carefully
Upstream].
Algemene Energie Raad AER (2005), Gas voor morgen, ISBN 90 74357 407 [Gas for Tomorrow]
Benders, R.M.J. (1996), Interactive simulation of Electricity Demand and Production, PhD Study
University of Groningen, ISBN 90-367-0631-9
Boer, B., (2004), Feasibility of a High Temperature Nuclear Reactor as Back-up System for Wind
Power, MSc Thesis, Report no. ET 2150, Delft University of Technology
BP (2007), BP Statistical Review of World Energy, June 2007
Boisseleau, F. (2004), The Role of Power Exchanges for the Creation of a Single European Electricity
Market: market design and market regulation, PhD study, TU Delft, DUP Science, ISBN 90-407-24733
CBS (2004), Duurzame energie in Nederland 2003 [Renewable Energy in the Netherlands 2003].
CBS (2005), Duurzame energie in Nederland 2004 [Renewable Energy in the Netherlands 2004].
CPB (2005), Werkgelegenheid en toegevoegde waarde per bedrijfstak, 2001-2020 en 2021-2040,
CPB Memorandum [Employment and Added Value per Business Sector 2001-2020 and 2021-2040]
DTE (2002), Beslissing op het bezwaar van de Federatie van Energiebedrijven in Nederland
(EnergieNed) gericht tegen het besluit van de directeur van DTE van 19 maart 2002 tot wijziging van
de Meetcode met betrekking tot het openbaar maken van meetgegevens, No. 100696/15 [Decision on
objection of EnergieNed against the decision of DTEs director of March 19, 2002, to modify the
Measuring code and make public of measuring data]
Delarue, E., Dhaeseleer, W. (2007), Adaptive mixed-integer programming unit commitment strategy
for determining the value of forecasting, Applied Energy 85 (2008) 171 - 181
ECN (2002), Bedrijfseconomische beoordeling van twee CO2-vrije opties voor electriciteitsproductie
voor de middellange termijn, Exploitatieverlenging kerncentrale Borssele en offshore windenergie,
ECN-C--02-055 [Economic assessment of two CO2-free options for the production of electricity in the
mid- to long-term: Extension of Borssele nuclear plant operations and offshore wind energy]
ECN (2005), Inzet van biomassa in centrales voor de opwekking van electriciteit, ECN-C--05-088
[Deployment of Biomass in Power Plants for the Generation of Electricity]
ECN (2005), Referentieramingen energie en emissies 20052020, ECN-C--05-018 [Reference
Projections for Energy and Emissions 20052020].
ECN (2005), The Next 50 years: four European energy futures, ECN-C--05-057.
Electro-Techniek (1964), Algemene aspecten van economische lastverdeling, 42e jaargang, No 24,
pp. 555-580 [General Aspects of the Distribution of the Economic Burden]
E.ON Netz (2005), Wind Report 2005
Essent (2004), Milieu overheidsjaarverslag 2003 Vestigingsplaats Geertruidenberg Amer en
Dongecentrale [Government Environmental Report for 2003: Geertruidenberg Amer and Donge power
plant sites]
Eurelectric (2007), The Role of Electricity, A New Path to Secure Competitive Energy in a CarbonConstrained World

227

References

European Commission EC (2001), Green Paper, Towards a European Strategy for the Security of
Energy Supply, ISBN 92-894-0319-5
European Union EU (2001), Richtlijn van het europees parlement en de raad van 27 september 2001
betreffende de bevordering van elektriciteisopwekking uit hernieuwbare energiebronnen op de interne
elektriciteitsmarkt, 2001/77/EG [Directive from the European Parliament and Council of September 27,
2001 on the promotion of generating electricity from renewable energy sources on the internal
electricity market]
European Union EU (2002), Council decision of 25 April 2002 concerning the approval, on behalf of
the European Community, of the Kyoto Protocol to the United Nations Framework Convention on
Climate Change and the joint fulfillment of commitments thereunder, 2002/358/CE
ExxonMobil (2004), The Outlook for Energy, A 2030 View
Frencken, H. (2003), Scenarios for the Future, Workshop ICLON, Leiden University.
Gasunie Trade and Supply B.V. (2006), Jaarverslag 2005 [Annual Report 2005]
Hiremath, R.B., Shikha, S., Ravindranath, N.H. (2005), Decentralized energy planning; modeling and
application a review, Renewable and Sustainable Energy Reviews, Volume 11, Issue 5, Junu 2007,
Pages 729 - 752
IPCC (2000),Emissions Scenarios, Summary for Policymakers, a special report of the IPCC Working
Group III
Jebaraj, S., Iniyan, S. (2004), A Review of Energy Models: renewable and sustainable energy reviews
10 (2006) 281-311
Kip, W.N. (2001), Antwoord aan EZ over energiecrisis in Californi, Vergadering Bestuur EnergieNed
d.d. 6 september 2001 [Answer to the Ministry of Economic Affairs about the energy crisis in California,
EnergieNed Board Meeting, September 6, 2001].
Landelijk Afvalbeheerplan 20022012 (2004), Deel 1 Beleidskader. Gewijzigde versie van april 2004
[National Waste Management Plan 2002-2012. Part 1, Policy Framework. Revised, April 2004].
Landelijk Afvalbeheerplan 20022012 (
2002-2012. Part 2, Sector Plans].

), Deel 2 Sectorplannen [National Waste Management Plan

Landelijk Afvalbeheerplan 20022012 (2004), Deel 3 Capaciteitsplannen. Gewijzigde versie van april
2004 [National Waste Management Plan 2002-2012. Part 3, Capacity Plans. Revised, April 2004].
Lapuerta, C., Moselle, B. (2001), Recommendations for the Dutch Electricity Market, The Brattle
Group, Ltd.
McNair (2000), All Possible Wars? Toward a Consensus View of the Future Security Environment,
20012025, Paper 63 Chapter 2: Estimates, Forecast, and Scenarios.
Ministry of Economic Affairs, the Netherlands (2000),Energie en samenleving in 2050, Nederland in
wereldbeelden [Energy and Society in 2050: the Netherlands in world views].
Ministry of Economic Affairs, the Netherlands (2005), Olie en Gas in Nederland, Jaarverslag 2004 en
prognose 20052014 [Oil and Gas in the Netherlands, Annual Report 2004 and Forecast 2005-2014].
Ministry of Economic Affairs, the Netherlands (2006), Olie en Gas in Nederland, Jaarverslag 2005 en
prognose 20062015 [Oil and Gas in the Netherlands, Annual Report 2005 and Forecast 2006-2015].
Ministry of Economic Affairs, the Netherlands (2007), speech by the Minister of Economic Affairs,
M.J.A. van der Hoeven, on the achievement of wind energy targets, June 15, 2007, Ridderzaal

228

References

NAM (2005), Diverse onderdelen verzorgd door NAM van cursus Gaswaardeketen , Energy Delta
Institute (EDI), Groningen [Several modules by NAM of the course Natural gas value chain]
Novem (2004),Statusdocument Bio-energie 2004 [Status Document Bioenergy 2004].
N.V. Samenwerkende elektriciteits-productiebedrijven, SEP (1990s), Wiskundig model ten behoeve
van optimale selectie en optimalisatie instellingen [Mathematical Model for Optimum Selection and
Optimization Institutions]
N.V. Samenwerkende elektriciteits-productiebedrijven, SEP (1996), Elektriciteitsplan 19972006
[Electricity Plan 1997-2006].
N.V. Samenwerkende elektriciteits-productiebedrijven, SEP (1996), Toelichting Elektriciteitsplan
19972006 [Explanatory Notes to the Electricity Plan 19972006].
Operation Simulation Associates (OSA), Manual PowrSym3
Operation Simulation Associates (OSA), Description of the PowrSym3TM Computer Model
Operation Simulation Associates (OSA) (2005), PowrSym3 Major Features
Projectgroep Transitie Biomassa (2003), Biomassa: de groene motor in transitie. Stand van zaken na
de tweede etappe [Biomass: Green Motor in Transition: state of affairs after the second stage].
SenterNovem (2005), Actieplan Biomassa. Statusdocument Bio-energie 2004 [Biomass Action Plan.
Status Document Bioenergy 2004].
SenterNovem (2002), Implementatie bio-energie in 2010. Perspectief per Biomassa Technologie
Combinatie. Openbare samenvatting [Implementation of Bioenergy in 2010. Perspective per Biomass
Technology Combination. Public Summary].
Shell (2002), People and Connections, Global Scenarios to 2020
Shell (2005), The Shell Global Scenarios to 2025, The Future Business Environment: trends, tradeoffs and choices
TenneT (2002), Capaciteitsplan 2003-2009 [Capacity Plan 2003-2009]
TenneT (2006), Rapport Monitoring Leveringszekerheid 2005 2013, MR 06-231 [Report on
Monitoring Security of Supply]
Ter Borg, R.W. (2005), Electricity Load Modeling using Computational Intelligence, PhD TU Delft,
DUP Science. ISBN 90-8559-118-X
UCI (2006), The Economics of Nuclear Power, Briefing Paper, 8 April 2006
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2002) Environmental behavior of the Dutch electricity
supply industy, Energie Techniek 7/8 July/August 2002
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2004), Trading between countries Improving the balance
between Environment, Economy and Security of Supply or power play? The EU Situation, paper
presented at 19th World Energy Congress, Sydney, Australia, September 5-9, 2004
Van Eck, T. (2007), A New Balance for the Energy Sector: no longer a puppet in the hands of
technology, public interests and market, ISBN
Voorspools, K. (2004), The Modeling of Large Electricity-Generation Systems with Applications in
Emission-Reduction Scenarios and Electricity Trade, PhD Study Catholic University of Leuven, ISBN
90-5682-501-1

229

References

Weingand, E. (1995), Future Research Methodologies: linking todays decisions with tomorrows
possibilities.
Wilkinson, L. (2004), How to Build Scenarios: planning for long fuse, big bang problems in an era of
uncertainty
Wright, A. (2004), A Social Constructionists Deconstruction of Royal Dutch Shells Scenario Planning
Process
Wood, A.J., Wollenberg B.F. (1996), Power Generation, Operation, and Control, 2nd Edition, Wiley,
New York, ISBN 0 471 58699 4.
Werkgroep Afvalregistratie (2004), Afval verwerking in Nederland: gegevens 2003 Vereniging
Afvalbedrijven [Waste processing in the Netherlands: figures for 2003 from the Dutch Waste
Management Association].
Vereniging Afvalbedrijven VAV (2004), Energie uit Afval, Statusrapportage 2004 [Energy from Waste,
Status Report 2004].
Verkaik, J.W. (1999), Evaluation of Two Gustiness Models for Exposure Correction Calculations,
Journal of Applied Meteorology Volume 39, No. 9, pp. 1613-1626, 2000.
Zuyi Li, Guangyu He, Xueqing Chen (2000), Network Flow Based Algorithm on Dynamic Optimal
Dispatch, Department of Electrical Engineering, Tsinghua University Beijing, 100084, China.

Internet Addresses
1.
2.
3.

230

http://www.eurelectric.org
http://www.grida.no/climate/ipcc/emission/index.htm
http://www.tennet.org/

Part 5

Input Model and Analysis of Scenario Outcomes

232

Chapter 9 Input Model for Developed Scenarios

Chapter 9
9

Input Model for the Developed Scenarios


9.1

Introduction

We have defined future scenarios in Chapter 7, which consequences must be converted into data that
serve as input for the model. The starting point is the Dutch electricity demand which will be analyzed
in Section 9.2 and resulting in an electricity load curve. This curve will be adapted for the analysis of
future scenarios. Section 9.3 describes the supply side of the individual modules whereby the
development of the share in total generation of the individual module over time and the expected midterm (2008 - 2012) and long-term developments (to 2020 - 2025) will be analyzed as starting points for
the future scenarios. This in combination with the financial parameters provides the demand, supply,
and financial input for the formulated future scenarios as described in Section 9.4.
9.2

Model Input: Demand Side

9.2.1

Introduction

To calculate supply and demand, an electricity load curve that is as close to reality as possible must
be created. The obtained electricity load curve should lead to the same electrical energy consumption
over time as the electricity balance that was applied to the selected scenario. Therefore the hourly
load values in the load curve can be adjusted with use of a percentage multiplier so that the
forecasted total electricity demand can be met for the selected scenario.
9.2.2

Electricity Demand

The Dutch electricity demand is shaped by consumption in four sectors: industry, transport, tertiary
and residential. Each sector has its own characteristic demand pattern. The development of the
expected economic growth, the gross domestic product (GDP), is an important factor when
determining the annual increase in electricity consumption. In Figure 9.1 it is shown that electricity
demand has been almost linearly proportional to the growth of GDP (see linear regression curve for
period 1996 - 2005). Only large variations in GDP are cancelled out.
In the CPBs scenario studies Four Futures of the Netherlands (2005) and Prosperity and
environment, a scenario study for the Netherlands in 2040 (2006), see also Section 7.2.3 at part
ECN, Reference Projections for Energy and Emissions 2005-2020, four scenarios and their effects
on GDP and employment are given. For all scenarios average an annual growth in GDP is expected in
range between 1.0 and 2.9% for 2002 - 2020 and 0.4 and 2.3% for 2021 - 2040. For the mid- and long
term electricity consumption we expect a continuation of the linear relation between annual GDP
growth and growth rate of electricity consumption for all scenarios.

233

Part 5

1999

4.0

1998
2000
3.5

1997
1996

3.0

Growth GDP [%]

2.5

2.0

2004
2005

1.5

2001

1.0

0.5
2002
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

-0.5
2003
-1.0

Growth Electricity Consumption [%]

Figure 9.1: Growth in GDP versus increase in electricity demand (Source: CBS, CBS Statline)
The Dutch electricity demand for one year can be illustrated by means of an electricity balance-sheet,
see Table 3.1, Chapter 3 for 2002 and 2003.
9.2.3

Electricity Load

To represent the demand side in the model, we take the electricity load expressed in MWe load at
hour i with i = 1 to 8760 h. This load varies constantly, but is characterized by a clear day/night pattern
and a winter/summer pattern. The current measurement data published by TenneT on their website
pursuant to Article 4.6 of the Dutch Measurement Code (DTE, 2007) is the basis for the created
demand curve. Three aspects of the published measurement data are important:
1. It refers to electricity that is fed into the grid, in other words: decentralized capacity is not
included in the published measurement data.
2. The measurement data relates to connected parties with an installed capacity of 10 MW or
more.
3. TenneT stated (~2005): The information contained in this publication is supplied by the grid
operators. The underlying data exchange was modified per 1 April 2004, from which date
onwards publication is based on data supplied by all of the grid operators (the data for one
grid operator was missing prior to this date). For this reason, a trend rupture may have been
created, making data starting 1 April 2004 unsuitable for comparison to earlier data. Because
this information is supplied by third parties, TenneT cannot vouch for its correctness, accuracy
or completeness.
Therefore the curve applied in the model (references year 2003) is derived from the curve for 2005.
The following procedure was followed to create a demand curve that could be used for the model:
1. Analyze measurement data published by TenneT.
2. Determine decentralized capacity that has to be added.
3. Estimate the < 10 MWe power fed into the public grid.

234

Chapter 9 Input Model for Developed Scenarios

Finally to create a realistic demand curve, the volume under the demand curve is made equal to the
total electricity demand (including grid loss) in the Netherlands (excluding export) as supplied by CBS
Statline.
Analysis of Measurement Data Published by TenneT
The measurement data that TenneT publishes per program time unit (PTU) is divided into:
The sum of measurement data supplied by grid operators for each connected party with a
generation capacity of 10 MW or more, providing it is fed into the grid, augmented by the total
energy exchanged in the cross-border grids (in MWh/PTU)
The sum of programmed imports (>0 in MWh/PTU)
The sum of programmed exports (<0, in MWh/PTU).
The PTU corresponds to electrical energy (MWh) in 15 minutes blocks. For the demand curve four
PTU blocks are summarized to create an hourly load and with 8760 hours a full year is created.
Share of Decentralized Power in the Demand Curve
Decentralized generation refers to electricity generation and consumption on an industrial or other
location it self and this electricity is not supplied to the public grid (see also Section 8.3.3). For
example, supplied to a locations own operating grid to power machines, lighting etc. Table 3.1 of
Chapter 3, shows that the share of decentralized power for 2003 is approximately 12 TWh. This,
however, does not establish what the load is at various times. Decentralized generation is fully
supplied by CHP units and the generated load is estimated for each CHP sub module (Industrial, DH
and Small CHP) based on CBS Statline and information received from market parties. The results of
the generated load of the sub module are summarized to a total decentralized load.
Power < 10 MWe Supplied to the Public Grid
Individual plants with a power of < 10 MWe are only found in Other Electricity Producers in the form of
CHP units (all types excluding DH CHP). First, the focus is on CHP power in the form of CCGT, GT +
HRSG and boiler + ST. The share of installed power < 10 MW amounts to 250 MWe of which between
75 and 100% is supplied decentralized, say some 200 MWe. In any case, it is clear that the remaining
50 MWe that is supplied to the public grid has no more than a marginal effect on the total demand
curve, which is why it is not included separately.
If we focus on gas-engine powered CHP, the total capacity of a unit is < 10 MWe. This decentrally
supplied power (see above) and the power supplied to the public grid have to be added to the TenneT
curve.
The other plants (< 10 MW) that generate electricity such as hydropower, and photovoltaic solar
energy, are often so small (1 MW level) that they cannot be included separately. Wind load is included
in the cluster (wind farms) data that the grid operators supply to TenneT.
The final demand curve applied in the model is displayed in Figure 9.2.

235

Part 5

20000

20000

15000

10000

18000

5000

0
0

20

40

60

80

100

120

140

160

180

16000

14000

Load [MWe]

12000

10000

8000

6000

4000

Load Ned
TenneT
CHP DC

2000

GasEng DC
GasEng Grid

0
0

1000

2000

3000

4000

5000

6000

7000

8000

9000

Figure 9.2: Created electricity demand curve for 2003 (Source: TenneT, CBS Statline, own information)

9.2.4

Heat Demand

The heat demand (process heat and district heat) that is related to the electricity supply can be broken
down into heat supplied by electricity production companies and by other electricity producers.
For the electricity production companies, data was taken from the annual environmental reports for the
individual plants that supply process heat or district heating, see Table 9.1. As input for the model the
annual pattern for heat demand for process heat in industrial projects is assumed to be constant. The
heat demand for district heating (in TJ) is distributed over the months using standard degreedays/year and is then distributed over the hours (see Section 9.2.5) according to a day and night
pattern of real DH projects. Forecasts for the expected growth in district heating are created in Section
9.4 for existing district heating areas for the horizons used in the scenarios.
City / Location

Type of generation

Process heat areas


Electrabel Nijmegen
IJmond Corus
Swentibold DSM
Electrabel Air Products
Shell Moerdijk
Total Heat

CHP / CONV
CHP / CCGT
CHP / CCGT
CHP / GT+HRSG
CHP / CCGT

District heat areas


Geertruidenberg
Almere
Amsterdam
Den Haag
Leiden
Purmerend

236

DH / CONV
DH / CCGT
DH / CCGT
DH / CCGT
DH / CCGT
DH / CCGT

Heat Supplied
in 2003 [TJ]

750
630
5,690
1,577
4,257
12,904
2,800
2,250
1,130
950
850
1,050

Chapter 9 Input Model for Developed Scenarios

Rotterdam
Utrecht
Total Heat

DH / CCGT
DH / CCGT

7,500
3,800
20,330

Total Process + District

33,234

Table 9.1: Heat demand (process heat and district heating) in areas of electricity production
companies (Source: Annual environmental reports individual generation units, own information)
For the other electricity producers, a combination of data from the annual environmental reports and
data from Statistics Netherlands (CBS) is used whereby the heat demand is related to the value
supplied by CBS for decentralized generation. As input in the model the heat demand for the clusters
that supply process heat are assumed to be constant and the heat demand for indoor heating (district
and greenhouse heat) are broken down over the months using standard degree-days. Table 9.2
provides an overview.
City / Location

Process heat
IND CHP
WASTE INCINERATOR
OTHERS
Total Heat
District heat
CCGT DH
WASTE INCINERATOR
NON FOSSIL
SMALL CHP
Total Heat
Total District + Process

Type of generation

CHP / CCGT
CHP / GT+HRSG
CHP / BOILER + ST
CHP / WI
CHP / OTHERS

DH / CCGT
DH / WI
DH / BMC
DH / GH / GASENGINE

Heat Supplied
in 2003 [TJ]

58,684
55,197
29,163
3,483
146
146,673
1,890
613
1,184
26,334
30,021
176,694

Table 9.2: Heat generation by other electricity producers (Source: CBS Statline, Annual environmental
reports individual generation units, own information)
In the process heat of industrial CHP CCGT, two large power plants are included, Elsta (475 MW)
supplying steam to DOW Chemical Europe in Terneuzen and Delesto (340 MWe) supplying steam to
Akzo Nobel. Total amount of process heat supply is 10,092 TJ to DOW and 4,275 TJ to Akzo Nobel.
The installed capacity of these installation account for approximately 30% of the total industrial CCGT
installations.
9.2.5

Heat Load

To represent the demand side in the model, we take the heat load expressed in MWth load at hour i
with i = 1 to 8760 h. The heat load patterns of process steam demand in industrial projects of CHP of
the electricity production companies are assumed to be constant. The heat load patterns of DH
projects of CHP of the electricity production companies varies with the seasons and day and night
pattern. The heat load patterns used are based on real measurement data of three larger DH projects
in the Netherlands (see Figure 9.3).

237

Therrmal Load [MWth]

Part 5

400

160

350

140

300

120

250

100

200

80

150

60

100

40

50

20

Utrecht
Amsterdam

0
0

1000

2000

3000

4000

5000

6000

7000

8000

0
9000

Purmerend

Figure 9.3: Measured heat demand curves for 2003 (Source: Nuon, Annual environmental reports
individual generation units)
The heat load for process steam and district heat for CHP and small scale CHP of the other electricity
producers, excluding the large power plants Elsta and Delesto, are displayed per month in the submodules.

9.3

Model Input: Supply Side

9.3.1

Introduction

Electricity demand and supply must always be in balance. The electricity demand curve that was
developed for the different scenarios must therefore always be in balance with the production capacity
of the following modules:

Electricity production companies


Other electricity producers
Import/export

The production capacity of these modules consists of different types of plants whose technical
systems are described in Part 2. The main purpose of this chapter is to describe the expected midterm (2008 - 2012) and long-term developments (to 2020 - 2025) of the production capacity in terms of
installed capacity, electricity produced and used technology. This will differ for the different scenarios,
and are based on several factors, which are model input parameters, that influence these
developments which are analyzed hereafter. For the electricity generating companies, a short analysis
of the past production capacity developments, known and unknown plans for new developments, the
expansion of existing locations, the re-commissioning of plants and demolishing plans are specified.
For other electricity producers, these developments are described by cluster. In addition, a short
analysis of the electricity supplied in the past is provided for each module and an attempt is made to
explain trends.

238

Chapter 9 Input Model for Developed Scenarios

9.3.2

Factors that Influence the developments of supply

For the modules mentioned in the introduction the following factors influence the production capacity:
Electricity demand in the Netherlands
Gas/Coal/Electricity price ratio
Price level of surrounding countries for import/export
Stimulation measures (MEP scheme)
Mandatory unit commitment.
Electricity Demand in the Netherlands
The electricity demand in the Netherlands has steadily increased since 1995 (see Figure 9.4).
Analysis of the CBS electricity balance shows that the share of supply from electricity generating
companies remained reasonably constant between 1995 and 1997, increased slightly in 1998, and
decreased strongly in 1999. Reason, in 1999, each market player was allowed to import. With the
price difference with surrounding countries as a driving force this, resulted in a substantial increase in
the amount of imported electricity. In 1999 alone, imported electricity increased by 56% compared with
1998. The supply from other electricity producers also started increasing in 1995 at the expense of the
supply from electricity generating companies. After the all-time low of 1999, the supply from electricity
generating companies has constantly increased again until 2004.
Noteworthy for the supply from other electricity producers is that after a period of constant growth
between 1995 and 2000, a continuous decrease took place starting 2001 and ending in 2004. This is
related to the commitment of CHP units on the deregulated electricity market since 2001 with its
negative spark spreads for off peak hours (see also Section 8.4.2). This forced installation owners of
CCGT and GT + HRSG based CHPs to run their units in part load or even shutdown during off peak
periods.
The share of supply from electricity production companies (the largest share!) is the balancing item
that satisfies the domestic electricity demand. In other words, the electricity demand after import and
after supply from other electricity producers determines the supply from electricity generating
companies. Supply from other electricity producers is primarily determined by other factors than just
the Dutch electricity demand.
120,000

Consumption + gridlosses
Elec.Prod Comp.

Import
free

Liberalized generation

Last "SEP"
year

Other Elec.Prod. Comp


Net Import
100,000

[GWhe]

80,000

60,000

40,000

20,000

0
1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

Figure 9.4: Overview of demand and supply and demand between 1995 and 2005 (Source: CBS
Statline)

239

Part 5

Gas/Coal/Electricity price ratio


As explained in Section 8.4.2, the electricity price for off peak periods is determined by the cheapest
generation option and this is depending on the fuel price ratio for coal and natural gas. The oil and
coal prices have developed as displayed in Figure 1.9 of Chapter 1. As can be seen, in 1999 the gas
price dropped significantly and came close to the coal prices and gas fired generation became the
cheapest option for off peak periods. The electricity generated by the coal-fired units in that year had
to decrease as much as possible on weekday nights, and even switched off on weekends in order to
avoid loss of revenues caused by a negative off peak spark spread.
Price Level of Surrounding Countries for Import/Export.
The import (or export) of electricity from surrounding countries as Germany, Belgium and France is
driven by the differences in base and peak load prices. In case the base load price differences are
small compared with peak load price differences, it means small price differences during the off peak
periods occur. Treat in electricity will then mainly take place for the peak periods and less for off peak
periods. This is further explained in Section 9.3.5.
Stimulation measures (MEP scheme)
The MEP was introduced on January 1, 2003 and first applied in the summer of 2003 (Ministry of
Economic Affairs, 2007) to encourage the production of electricity generated with sustainable sources
and CHP. For CHP, the MEP scheme was replaced by the CO2 Indexing Scheme as of July 1, 2004
(see also Van Eck, 2007). In this scheme, see Figure 9.5, the amount of "CO2 neutral or blue
electricity generated by a CHP is determined whereby this amount depends on the plant's energetic
performance. The plants actual electricity production, measured at the generator, and net supply of
usable heat on a monthly basis are converted to a quantity of fuel based on the reference efficiencies
for power and heat. That quantity less the combined heat and powers plant real fuel consumption
represents the energy saved. That saving and the reference electricity efficiency are then used to
calculate the amount of so-called CO2 free kWhs. The correct heat/power ratio with a high total
efficiency (electric + thermal) results in the highest share of blue electricity. The compensation for
every CO2 neutral MWh electricity generated, which for the second half of 2004 was set at 26 /MWh,
is added to the market price. The compensation is restricted because of EU rules and cannot exceed
50% of the difference between cost and market prices. The scheme covers only the first 1,000 GWh of
a units annual output and is not a permanent environmental incentive scheme. After December 31,
2005, only CHP projects with an age < 10 year, receive the compensation.

Figure 9.5: MEP CO2 Indexing Schema (Source: Cogen Netherlands)


240

Chapter 9 Input Model for Developed Scenarios

For new renewable capacity there is a guaranteed rebate over ten years depending on the type of
installed capacity (wind on- or offshore, hydro or solar). The rebate for requests for new renewable
wind on-shore, hydro and solar projects and received after August 18, 2006 was set at 0 /MWh by
the Minister of economics affairs (EnerQ, 2007). For wind offshore this was already the case on May,
10, 2005.
For waste incinerators two criteria must be fulfilled to be considered for the MEP scheme (Ministry of
Economic Affairs, 2005):
1. An energetic efficiency of at least 22% consisting of the share of electricity supplied to the
electricity grid, plus two-thirds of the heat supplied to third parties (e + 2/3th).
2. Only new waste incinerators or expansions in incineration lines with an energetic efficiency
higher then 22% receive the rebate.
The rebate starts with a low tariff of 3/MWh for a Waste Incinerator with an energetic efficiency of
22% and increase to a value of 38 /MWh in case the energetic efficiency is lager then 30%.
For biomass incineration or fermentation the rebate is depending on the size, > 50 MW (central co
firing), > 10 MW or 50 MW, 10 MW and if the biomass is pure or mixed biomass. The rebate for
requests for new biomass projects > 50 MW and received after May 10, 2005 was set at 0 /MWh by
the minister of economics affairs. For new biomass projects 50 MW this was the case on 18 august
2006, however for requests received before this date the rebate is guaranteed over ten years.
In Table 9.3 an overview is given for the applied MEP rebate for the period 2003 - 2007.
MEP - category - production period (/MWh)
CHP > 120 MWe
CHP <= 120 MWe
CHP gas engines
Wind Onshore
Wind Offshore
Hydro
Solar
Bio-energy
Waste incineration < 26% efficiency
Waste incineration >= 26% efficiency
Biomass incineration
Mixed biomass > 50 MW
Pure biomass
> 50 MW
Mixed biomass <= 50 MW
Pure biomass <= 50 MW
Biomass fementation
Landfill gas
Sewage treatment plants

2003-II

2004-I

2004-II

2005-I

2005-II

2006-I

2006-II

2007

5.7
5.7
5.7

5.7
5.7
5.7

26
26
26

22
22
22

22
22
22

18
33
33

18
33
33

0
21
22

49
68
68
68

48
67
67
67

63
82
82
82

77
97
97
97

77
97
97
97

77
97
97
97

65
97
97
97

65
97
97
97

0
29

0
29

0
29

0
29

0
29

0
29

3-9
11 - 38

3-9
11 - 38

29
48
29
68

29
40
29
67

29
55
29
82

29
70
29
97

29
70
29
97

29
70
25
97

25
25
25
97

25
25
25
97

21

21

21

13

13

Table 9.3: Overview MEP rebate for the period 2003 2007 (I =jan- jun, II = jul - dec)(Source: EnerQ,
2007)
In 2008 the MEP scheme will be replaced by the stimulation regulation renewable energy production
(SDE) scheme for sustainable sources and CHP.
Mandatory Unit Commitment
A number of units must be committed (must run) independent of electricity market conditions. In the
utilities situation as well as on the free market, this primarily concerns:
Blast furnace gas units
CHP units
DH units
Small CHP.
Blast furnace gas (BFG) comes available from Corus in large quantities and this gas has to be burned.
In Velzen at least 1 of the 2 conventional units (VN24 or VN25) is in operation in order to burn the
BFG. IJmond is a CHP unit burning BFG and is always in operation if technically available.

241

Part 5

The heat demand of CHP units is dictated by the industrial processing plant en is normally constant
(24 hours/365 days) also during the night and weekends (off peak). The CCGT, gas turbine and heat
recovery steam generator (GT + HRSG) units can be powered down during nightly off peak periods
and on weekends if the variable costs exceed the income (see Section 8.4.2). A complete shutdown
during these periods however is not an option because other considerations such as the security of
supply of heat production and the risk of not being able to startup the plant up again due to failures
play an important role besides electricity market conditions.
For large DH units supplying heat for district heat areas the operation depends on the heat demand
and the number of individual DH units at one location. For example, in Utrecht, DH can be supplied by
Lageweide 6 as well as Merwedekanaal 10, 11 or 12. At least one unit will be in operation in the
summer period and two in the winter period. The third unit is committed when one of the two is not
available or in case of cold weather conditions which requires extra heat capacity. For locations that
have only one unit with one gas turbine, the unit, if technical available, is committed at full or partial
load depending on the dispatch. For plants that have two gas turbines, one of the gas turbines can be
shut down when there is a low heat load, while the other can be run at full or partial load to cover the
heat demand. In case heat demand is low in the summer period and heat boilers and/or buffers are
available DH units can shutdown completely and there is no real mandatory unit commitment.
The largest share small scale CHP i.e. gas engines can be found in greenhouse farming (see also
Section 4.5.5) where they are used to supply electricity, heat, and at a few locations, CO2. The heat
demand in greenhouse farming and other sectors that use gas engines, such as the heat distribution
of energy companies, is seasonal. This creates large discrepancies between the heat demand during
the winter and summer months. Due to the low heat demand in the summer, the number of operating
hours is low, which means that the demand can only be fulfilled in combination with heat buffering.
More operating hours are possible in the winter. Compared with industrial CHP, the advantage of gasengine powered CHP is that a gas engine can be easily switched on and off. Heat can be buffered
because gas engines are used primarily to supply hot water and are hardly used to supply steam.
Greenhouse farmers and heat distribution projects buffer heat. During the off peak period, reducing
the capacity is equivalent to switching off the engine. A gas engine is easily started and stopped,
which is why there is no real mandatory unit commitment compared with CHP and DH units.
9.3.3

Electricity Production Companies

The installed electrical capacity is broken down as follows (see Figure 9.6):
5,000
4,500

3,960

3,500
3,000
2,308

2,500

2,100
1,826

2,000

1,753

1,500
759

1,000

449

IND CHP

CCGT DH

IGCC

NUCLEAR

HW CCGT

CONV GAS

CONV COAL

253

168

CCGT

500

PEAK

Installed Capacity [MWe]

4,000

Figure 9.6: Electricity production companies: installed capacity 2003 (Source: CBS Statline, own
Information)

242

Chapter 9 Input Model for Developed Scenarios

General
With a share of almost 50%, the contribution of CCGT based generation capacity is highest followed
by 45% conventional generation capacity. The remaining capacity consists of gas turbines units for
peak load, one nuclear unit in Borssele and one IGCC unit in Buggenum.
A number of smaller DH plants, namely Enschede, Erica, Klazinaveen and Helmond (2) are owned by
a electricity production company but are not included in this module. The commitment for these plants
is determined outside the PowrSym3TM model and falls therefore under Supply by Other Electricity
Producers, the commitment of which is determined by separate sub modules (see Section 8.4.2).
Development of the Electricity Supply over Time
Figure 9.7, which displays the supply from electricity production companies, was created from data in
CBS's electricity balance 2000 - 2004. CBS does not state the units for which the gross and net
generated electricity is included, so we carried out our own research to determine which of the units
owned by electricity production companies CBS most likely included in its report. We studied the
environmental reports of as many individual units/locations as possible to obtain a clear picture of the
net electricity generated by the units over the years. The discrepancy between CBS and our own
analysis for the period 2000/2004 is between -2.0% and +2.7%. During and interview with CBS in
2006 our analysis was discussed and compared with the CBS data. It became obvious that CBS used
the supply data for all of the production units that fell under the SEP scheme that applied until 2000
and this is according to our own analysis. Figure 9.6 displays the breakdown of the share of DH, CHP,
and the other units we obtained from our own analysis.

70,000

60,000

Electricity Poduction [GWhe]

50,000

40,000

CBS
TOTAL UNITS
DH CCGT
CHP CCGT

30,000

OTHERS

20,000

10,000

0
1999

2000

2001

2002

2003

2004

Figure 9.7: Developments in the electricity supply from electricity production companies (Source: CBS
Statline, own information)
Mid-term Developments (2008 - 2012)
The mid-term developments in the share of production capacity of the electricity production companies
are, on the one hand, a function of the age structure of the current units and their corresponding
refurbishment or re-commissioning plans, and on the other hand a function of the plans for new units.
The age structure of the electricity production companies units is displayed in Figure 9.8 (not including
Lageweide 5 (demolished) and Flevo 30 (mothballed)).

243

Part 5

3,000

20,000

10 years
20 years
30 years

16,000

Installed Capacity [MWe]

2,071

14,000

2,000
1,766

12,000
1,537

1,500

10,000

1,334

1,244
8,000

1,078

997

1,000

6,000

293

233
0
2002 < L 2004

1998 < L 2000

1996 < L 1998

1994 < L 1996

1992 < L 1994

1990 < L 1992

1988 < L 1990

1986 < L 1988

1984 < L 1986

1982 < L 1984

1980 < L 1982

1978 < L 1980

1976 < L 1978

1974 < L 1976

1972 < L 1974

2,000
43

2004 < L 2006

103

48
1970 < L 1972

4,000

483

442

2000 < L 2002

479

500

Cumulative Capacity [MWe]

2,500

18,000
2,422

Installed
Cumulative

Figure 9.8: Age structure of electricity production companies generation units (Source: Annual
environmental reports individual generation units, own information)
SEP had set the original lifetime for decommissioning at 25 years (TenneT, 2002) regardless of the
unit's condition. Now that the electricity sector has been deregulated, other factors play a role in the
consideration whether to decommission a plant or not. There is (status 2002 - 2004) some 4,000 MWe
of generation capacity that is older than 25 years and this figure is going to rise in the coming years.
However, with today's modern maintenance and inspection levels, and the possibilities of extending
the lifespan of existing units, life expectancy of units can be extended with 50% (TenneT, 2002), and
therefore it is unlikely that much capacity from the electricity generating companies will have to be
decommissioned in 2008 - 2012. The assumption is that all of the capacity with an age of > 35 years
between 2005 and 2015 will be decommissioned. This means units that were commissioned between
1970 and 1980 will no longer be operational except for the units for which extensive modernization
plans are carried out. Nowadays 3 HW CCGT plants (Bergum 10, 20, Eems 20) are modernized to
extend lifetime until a period between 2010 and 2015, see Table 9.3. For the scenario study however
they are taken out of operation in 2012, in short, this means:
HW CCGT:
Total

1,360 MWe decommissioned (BG10,11, EC20)


1,360 MWe decommissioned

The known plans for new power plants and modernization plans for the mid-term (situation as of 2007)
are given in Table 9.4 which provides and overview:

244

Chapter 9 Input Model for Developed Scenarios

Generator

Type

Completed
E.ON Benelux, UCML
Intergen, Rijnmond
Electrabel, Flevo 30 GT

CHP/CCGT
CHP/CCGT
PEAK

Planned:
E.ON, Maasvlakte
Electrabel, Maasvlakte
Essent, Borssele,Geertruidb
RWE, Eemshaven
Nuon, Eemshaven
Delta/RWE, Sloe
Electrabel, Flevo
Enecogen Europoort

CONV
CONV
CONV
CONV
IGCC
CCGT
CCGT
CCGT

Electrical output

Fuels

Commissione
d

77 MWe
800 MWe
120 MWe
997 MWe

Natural Gas
Natural Gas
Natural Gas

End 2003
2004
2004

1,100 MWe
600 - 800 MWe
800 - 1,100 MWe
1,600 MWe
1,200 MWe
820 MWe
800 - 900 MWe
840 MWe

Coal / Biomass
Coal / Biomass
Coal / Biomass
Coal / Biomass
Multi-fuel
Natural Gas
Natural Gas
Natural Gas

2011
2011 - 2012
2013
2012
2011
2009
2008/2009
2008 - 2009

7,760- 8,360
MWe
Modernization
Electrabel, Gelderland 13
Electrabel, Bergum
Electrabel, Harculo
Electrabel, Eems 20
Essent, Amer 8
Essent, Claus plant,
overhaul
Essent, Claus plant,
expansion
E.on, Leiden, repl. GTs
E.on, The Hague, repl. GTs
E.on Maasvlakte
EPZ, Borssele
Recommissioning plans
Essent, Amer 7

COAL
HW-CCGT
HW-CCGT
HW-CCGT
CONV
CONV

603 MWe
664 MWe
350 MWe
695 MWe
645 MWe
640 MWe

Coal
Natural Gas
Natural Gas
Natural Gas
Coal
Natural Gas

To 2017
To 2012
To 2013
To 2012
To 2016
To 2015

CONV

920 MWe

Natural Gas

To 2034

84 MWe
112 MWe
1,040 MWe
479 MWe
Expansion 347
MWe

Natural Gas
Natural Gas
Coal
Uranium

To 2030
To 2030
To 2022
To 2034

DH/CCGT
DH/CCGT
CONV
NUCLEAR

CONV

400 MWe
Natural Gas
400 MWe
Table 9.4: Mid-term developments (Source: ANP Pers Support, Energeia, elektrabel.nl, eonbenelux.com, nuon.nl, utilities, own information, e.o)

2006

If all of the plans are carried out, the electricity production companies will have expanded their total
capacity by approximately 8,000 8,700 MWe (including modernization plans) compared with January
1, 2003. It is pretty certain that the modernization plans will be carried out. However, it is not certain
that all the new development plans will be carried out. Only the E.on coal/biomass conventional plant
is certain, and during the summer of 2007 the work is started for the building of the Sloe power plant.
Electrabels and Nuon's new development plans are in the study / permission phase. There are plans
to build new coal fired units but the long preparation and construction times do not enable them to be
all included in the first horizon. Nuclear power is being seriously considered (EZ, 2008) but has a very
long lead on time (see also Section 4.9, Chapter 4).
In order to carry out the scenario studies for the first sight period 2012, 23 units on top of the existing
units have been modeled for PowrSym3. These units can be applied for the above described known
plans for new developments but also for the unknown plans which are a function of the scenario. In
Table 9.5 an overview is given of these units and in appendix D the heat rate curves are displayed.

245

Part 5

First Sight Period 2012

Name

Conventional
Conventional
IGCC
IGCC
CCGT
CCGT
PBMR

Unit
Unit
Unit
Unit
Unit
Unit
Unit

1, Unit 2, Unit 3
4
5, Unit 6, Unit 7, Unit 8
9, Unit 10
11, Unit 12, Unit 13, Unit 14, Unit 15, Unit 16
17, Unit 18, Unit 19
20, Unit 21, Unit 22, Unit 23

Fuel

Unit Size (MWe)

Coal / Biomass
Multi Fuel
Coal / Biomass
Multi Fuel
Gas
Gas
Uranium

1,200
1,200
550
900
480
832
170

Number Total Capacity (MWe)


3
1
4
2
6
3
4

3,600
1,200
2,200
1,800
2,880
2,496
680

23

14,856

Table 9.5: Units built in PowrSym3 for first sight period 2012
Long-term Developments (2020 - 2025)
With the assumption that all of the capacity with an age of > 35 years between 2015 and 2025 will be
decommissioned, units commissioned between 1980 and 1990 will no longer be operational. Excluded
are units for which extensive modernization plans are carried out to extend lifetime until the period
after 2025, see Table 9.4. In short, this means:
Conventional coal:
Conventional gas:
HW CCGT:
Peak:
CCGT:
DH CCGT:
Total

2,730 MWe decommissioned (G13, MV1, MV2, A-81, BS12)


1,460 MWe decommissioned (CC (1 plant), VN24, VN25)
949 MWe decommissioned (HC60, HW7)
168 MWe decommissioned (all units)
120 MWe decommissioned (DNG)
380 MWe decommissioned (AL-1, ROC1, ROC2, MK10, MK11, PU-1)
5,807 MWe decommissioned

In order to carry out the scenario studies for the second sight period 2025, 30 Units (on top of the first
mentioned 23 units and existing units) have been modeled for PowrSym3. These units can be applied
for new developments which are a function of the scenario. For conventional, IGCC and CCGT plants
also the option of CO2 capture can be modeled with the assumption that these technology will be
available round 2025 against reasonable cost. In Table 9.6 an overview is given of the units and in
appendix D the heat rate curves are displayed.
Second Sight Period 2025 Name
Conventional
Conventional
IGCC
IGCC
IGCC
CCGT
CCGT
AWBR
EPR
PBMR

Unit
Unit
Unit
Unit
Unit
Unit
Unit
Unit
Unit
Unit

24,
28,
30,
36
37,
39,
43,
45,
47,
49,

Unit 25, Unit 26, Unit 27


Unit 29
Unit 31, Unit 32, Unit 33, Unit 34, Unit 35
Unit
Unit
Unit
Unit
Unit
Unit

38
40, Unit 41, Unit 42
44
46
48
50, Unit 51, Unit 52, Unit 53

Fuel

Unit Size (MWe)

Coal / Biomass
Multi Fuel
Coal / Biomass
Coal / Biomass
Multi Fuel
Gas
Gas
Uranium
Uranium
Uranium

1,200
1,200
600
1,200
1,200
500
1,000
1,465
1,600
300

Number Total Capacity (MWe)


4
2
6
1
2
4
2
2
2
5

4,800
2,400
3,600
1,200
2,400
2,000
2,000
2,930
3,200
1,500

30

26,030

Table 9.6: Units built in PowrSym3 for second sight period 2025
In total for 2012 and 2025 this means that 53 new units with a total installed power of approximately
41,000 MWe have been built for PowrSym3 and are available for the model. The choice of which units
are applied in the model depends, however, on the chosen scenario.

246

Chapter 9 Input Model for Developed Scenarios

9.3.4

Other Electricity Producers

The electrical installed power by the other electricity producers can be broken down as follows (see
Figure 9.9):

4,000

3,557

3,500

Installed Capacity [MWe]

3,000
2,500
1,513

2,000

1,016

1,500
1,000

414
247

500

OTHERS

WASTE INC

NON FOSSIL

SMALL CHP

CCGT DH

IND CHP

67

Figure 9.9: Installed power 2003 by other electricity producers (Source: CBS Statline, Cogen, own
information)
With a share of almost 52%, the contribution of industrial CHP based generation capacity is highest
followed by 22% gas engine capacity and 15% non fossil (of which 90% is wind capacity). Remaining
capacity consists of DH CCGT, waste incinerators and others.
Industrial CHP
Development of the Share of Supply over Time
Figure 9.10, which displays the supply from industrial heat/power, was generated from data in CBS
Statline with separate electricity production divided in clusters 1999 - 2005.

247

Part 5

16,000

14,000

Electricity Production [GWh]

12,000

10,000

8,000

6,000

4,000

CCGT

2,000

GT + HRSG
BOILER + ST

0
1999

2000

2001

2002

2003

2004

2005

Figure 9.10: Developments in the electricity supply from industrial heat/power (Source: CBS Statline)
As can be seen in Figure 9.10 the share of CCGT has only increased significant in 2001 compared
with 2000, while the share of GT+HRSG has decreased over the years and the share of Boiler + ST is
relatively constant. The increase of share of CCGT can be explained by the expansion of installed
capacity in 2001, see Figure 9.11 and the decrease of share of GT+HRSG is explained mainly by
decrease in operation hours. In fact operation hours decreased for all forms of industrial heat/power in
the period 1999 - 2005.
8,000

2,500

7,500

7,000

2,000

6,000

1,500

5,500

5,000

1,000

Operation Hours [h]

Installed Capacity [MWe]

6,500

4,500

4,000

500

CCGT
GT + HRSG
BOILER + ST

3,500

CCGT
GT + HRSG

3,000
1999

2000

2001

2002

2003

2004

BOILER + ST

2005

Figure 9.11: Installed capacity and operation times CHP units over the years (Source: CBS Statline)

248

Chapter 9 Input Model for Developed Scenarios

Mid-term (2008 - 2012) / Long-term (2020 - 2025) Developments


The possible mid and long-term heat/power developments are described by ECN (2005) in its report
WKK in de referentieramingen 20052020 (CHP in the reference projections 20052020). A
development is also described in the ECN (2005) report Referentieramingen energie en emissies
20052020 (Reference projections for energy and emissions 20052020). Both reports are used to
describe the heat/power developments, but the first report is the most important because it describes
the development of decentralized CHP with the same distinction between central and decentralized
distribution that is used in this study. Table 9.7 shows some of the key CHP data that is included in
ECNs Global Economy (GE) scenario (see also Section 7.2.3, Chapter 7).

Heat demand suitable for CHP


Covered by CHP
Thermal power CHP
Electrical power CHP
Savings
CO2 reduction

[PJ]
[PJ]
[GW]
[GW]
[PJ]
[Mton]

2003

2010

2020

839
190
9
5
46
3,8

877
204
11
7
36
3,1

945
201
11
9
41
3,6

Table 9.7: Key CHP data for GE scenario (Source: ECN, 2005)
The table shows that CHP is not used to its full potential. Of the heat demand specified, 35 - 40% is
used for direct under firing, for example, in oil refineries (heating crude oil, naphtha cracking). Using
CHP requires higher investments because the plants usually have to be built from scratch, which is
not likely to happen. The growth in CHP capacity is mainly a result of the growth in industrial steam
demand. The heat generated by small scale CHP plants (gas engines) for greenhouse farming and
service sectors is decreasing as a result of the lower heat demand in these sectors, whereas the use
of small CHP for illuminated cultivation is increasing. Neither the savings nor the CO2 reduction are
increasing because the reference technology for electricity generation is continuously improving.
As discussed by Van Eck (2007), there are a number of policy instruments to support CHP, such as
the MEP scheme, the EIA scheme (energy investment allowance), and CO2 emission trade. Table 9.8
shows the influence of the described policy instruments on the growth of CHP capacity according to
ECN (2005).
Electrical power
[GWe]

GE
Policy free as of 2005 (excl.
MEP, Emission trading and EIA)
GE, with extra 30 /ton CO2
GE, with double EIA

Thermal power
[GWth]

H/P ratio
[-]

2003

2010

2020

2003

2010

2020

2003

2010

2020

5,0
5,0

6,7
6,0

8,6
6,7

9,4
9,4

10,5
9,9

10,8
9,3

1,9
1,9

1,6
1,7

1,3
1,4

5,0
5,0

7,7
7,5

10,4
10,0

9,4
9,4

11,8
11,5

12,8
12,4

1,9
1,9

1,5
1,5

1,2
1,2

Table 9.8: Influence of policy instruments on the growth of CHP capacity (Source: ECN, 2005)
Without CHP policy, CHP will hardly grow. If todays policy is continued, the installed decentralized
CHP capacity is expected to grow to approximately 9 GWe. The heat/power ratio will decrease further,
which shows that more and more future plants will be expanded to generate electricity.
Besides possible CHP growth, also the modernization of existing units is of importance. In most units
(CCGT and GT+HRSG) only gas turbines have to be replaced by a newer or modified one in order to
increase electrical efficiency and to avoid an increase in electricity generation cost to much compared
with the new build power plants of the electricity production companies. In Figure 9.12 and 9.13 the
age of CCGT and GT+HRSG units is given.

249

250

0
49

22

2004 < L 2006

20
0
0

2002 < L 2004


2004 < L 2006

20 years
647

600

466
2,000

1,750

400
340
1,500

300
1,250

224
1,000

200

93

47

10 years

20 years

159

136

120
600

93
500

80
67
400

300

40

8
18

Cumulative Capacity [MWe]

10 years

Cumulative Capacity [MWe]

2000 < L 2002

1998 < L 2000

1996 < L 1998

2002 < L 2004

47

2000 < L 2002

98

1998 < L 2000

200

1996 < L 1998

221

1994 < L 1996

800

1994 < L 1996

60
1992 < L 1994

Installed Capacity [MWe]


500

1992 < L 1994

140
1990 < L 1992

93

1990 < L 1992

23
1988 < L 1990

1986 < L 1988

11

1988 < L 1990

180

1986 < L 1988

160
1984 < L 1986

1982 < L 1984

1980 < L 1982

1978 < L 1980

1984 < L 1986

100

1982 < L 1984

1974 < L 1976


105

1980 < L 1982

1976 < L 1978

1972 < L 1974

100

1978 < L 1980

1976 < L 1978

1974 < L 1976

1970 < L 1972

1972 < L 1974

1970 < L 1972

Installed Capacity [MWe]

Part 5

3,000

700
2,750

2,500

2,250

750

500

250

0
Installed

Cumulative

Figure 9.12: Age structure of CCGT of other electricity producers (Source: Cogen, own information)

1,000

900

800

700

43
200

100

Installed

Cumulative

Figure 9.13: Age structure of GT+HRSG of other electricity producers (Source: Cogen, own
information)

In the nineties most CCGT units were build and they are now approximately 10 years old. The
GT+HRSG units are somewhat older: 15 - 20 years.

Chapter 9 Input Model for Developed Scenarios

CCGT DH
There is no growth expected in this already small amount of installed capacity, actually two plants
have stopped supplying heat to greenhouse farmers because of the low electricity prices in the off
peak period. The medium and long-term developments are the same as described for industrial CHP.
Small CHP (Gas engines)
Development of the Share of Supply over Time
Figure 9.14, which displays the supply from small CHP, was generated from data in CBS Statline with
separate electricity production divided in clusters 1999 - 2005.

6,000

Electricity Production [GWhe]

5,000

4,000

3,000

2,000

1,000

0
1999

2000

2001

2002

2003

2004

2005

Figure 9.14: Developments in the electricity supply from small CHP (gas engines) (Source: CBS
Statline)
The supply from gas-engine powered CHP in the electricity supply to the grid has decreased until
2004. In 2005 the share has increased slightly because of the extra installed capacity in greenhouses,
see Figure 9.15. Operating hours however have decreased sharply as of 2004 which also means less
heat production from the gas engine.

251

Part 5

4,000

2,000

3,750

3,500

1,000

3,250

Operation Hours [h]

Installed Capacity [MWe]

1,500

500
3,000

Capacity
Hours
0

2,750
1999

2000

2001

2002

2003

2004

2005

Figure 9.15 : Developments of installed capacity and operation hours of gas engines heat/power
(Source: CBS Statline)
Mid-term (2008 - 2012) / Long-term (2020 - 2025) Developments
Gas engines are replaced after approximately 15 years of operation by a completely modified or new
gas engine in order to increase electrical efficiency. In Figure 9.16 the age of gas engines installed
power is given. For the expected mid-term and long-term developments see the description in the
section on industrial CHP.
3,000

700
10 years
601
600

2,500
500

Installed Capacity [MWe]

2,000
400
344
282

300

1,500

225
200

162

147

1,000

Cumulative Capacity [MWe]

436

-43

1998 < L 2000

2000 < L 2002

100
500

2004 < L 2006

2002 < L 2004

1996 < L 1998

1994 < L 1996

1992 < L 1994

1990 < L 1992

-100

1980 < L 1990

0
Installed
Cumulative

Figure 9.16: Age structure of gas engines of other electricity producers (Source: Own information)

252

Chapter 9 Input Model for Developed Scenarios

Renewable energy
Development of the Share of Supply over Time
The share of supply from renewable sources from 2000 to 2005 is displayed in Table 9.9.
2000

2001

2002

2003

2004

2005

GWh

GWh

GWh

GWh

GWh

GWh

142
829
8
p.m.
-

117
825
13
p.m.
-

110
946
17
p.m.
-

72
1318
31
p.m.
-

95
1867
33
p.m.
-

88
2067
34
p.m.
-

979

955

1073

1421

1995

2189

Bio-energy
o.w. Waste incineration
Biomass incineration
o.w.
Central co-firing
Decentral
Biomass fermentation
o.w.
Landfill gas
Sewage treatment plants
Others

1694
1003
414

2037
962
784

2556
942
1298

2225
959
962

2968
931
1756

4823
1001
3545

198
216
277

563
221
291

1082
216
316

757
205
304

1539
217
281

3310
235
277

153
108
16

160
115
16

176
119
21

166
111
27

134
126
21

127
119
31

Total (renewable)

2673

2992

3629

3646

4963

7012

Total (excl. waste inc. & co-firing)

1472

1467

1605

1930

2493

2701

Hydropower
Wind power
Photovoltaic solar energy
Thermal solar energy
Heat pumps
Heat/Cold storage
Subtotaal

Table 9.9: Developments in the share of renewable electricity (Source: CBS, 2007 )
The development in the share of renewable electricity is determined by:
The share of wind electricity from wind turbines, which strongly increased after 2002 in
particular because of the number of large wind turbines that were added after 2002 (see also
Figure 4.30 of Chapter 4).
Hydropower, which is dominated by three plants in the big rivers (share more than 90% of the
capacity). Because no new hydropower plants have been built since 1990, the annual
variability in electricity generation is caused in particular by the variability in water supply in the
big rivers. Little energy was produced in 2003 because it was a dry year along the Rhine and
Maas basins.
The installed capacity of photovoltaic solar energy in 2003 which almost doubled compared
with 2002 as a result of the energy incentive scheme (EPR), which enabled households to buy
solar panels for little money. This scheme was discontinued in October 2003. The total
contribution toward renewable electricity generation is only approximately 0.3% of the
domestic electricity consumption.
Bio-energy
o Central co-firing: The share of central co-firing increased sharply in 2002 compared
with 2001, but in 2003 this share sharply declined mainly because of the minimum
financial support from the government. The regulatory energy tax (REB) was

253

Part 5

decreased on January 1, 2003 and the substitution MEP scheme did not come into
effect until July 2003.
o Decentral generation, dominated by BMC in Cuijk, has slightly increased in 2005
compared with previous years.
Biomass fermentation
o Landfill gas: The share of generation from landfill gas has remained fairly stable in the
last years.
o Sewage treatment plants: The share of generation from sewage treatment plants has
remained fairly stable in the last years. Due to the drought in 2003, slightly less
electric energy was generated than in 2002 because less rainwater sludge was
processed.
o Other biogas: This mainly concerns biogas that is generated through the anaerobic
treatment of sewage and is used to generate electricity and/or process heat in the
food and paper industries. The share of generation from other biogas has remained
fairly stable in the last years.

Medium to Long-term Developments in Energy from Natural Sources (2012 - 2025)


The primary focus is on wind, hydropower, solar, and central and decentralized biomass incineration.
No significant developments are expected in the area of biomass fermentation.
Onshore Wind
Literature research conducted by ECN (2004) shows that the Netherlands has a realizable potential of
onshore wind power between 1.5 and 3.2 GW for the period up to 2050. IEA (2001) states the Dutch
government's goal, which is 1,500 MW in 2010 with an energy capacity factor of 26.6%. According to
ECN, these 1,500 MW are a realistic goal, but its achievement depends in part on government
incentives (MEP scheme).
Offshore Wind
According to ECN (2004), the realizable potential of offshore wind power is between 6 and 30 GW for
the period up to 2050. Here too, IEA (2001) uses the Dutch governments goal of 6,000 MW by the
year 2020 as its starting point. CA (2001) estimates a potential of 10,000 MW for the period 2000 2020. ECN considers this figure to be realistic. According to Lako et al (1998), the energy capacity
factor is 33.8% for near shore farms and 36.5% for offshore farms. CA (2001) assumes the factor will
be 37.9% in 2050. The application of offshore wind power also depends on government incentives.
(SDE scheme).
Hydropower
The limited amount of drop in Dutch rivers and the role that the rivers play in undisturbed drainage and
shipping reduce the potential of hydropower. SEP's E-plan (1996) assumes a potential of almost 0.
ECN (2004), however, assumes a potential for small hydro of 17 MWe in the low investment range.
Photovoltaic Solar Energy
To estimate the development of photovoltaic solar energy systems, SEP's E-plan (1996) uses the
estimates in the National Research Programme Solar Energy (NOZ-PV). This source of power was
clearly assumed to grow as much as 85 MW in 2006, with a continued growth to 250 MW in 2012.
Table 9.10 shows ECN's (2004) view of the realizable potential for the Netherlands.

254

Chapter 9 Input Model for Developed Scenarios

Table 9.10: Realizable potential for the Netherlands (Source: ECN, 2004)
The table shows that the number of systems will increase in particular between 2010 and 2020. This
has to do with the technological development of photovoltaic solar energy systems, particularly in
combination with the expected decrease of costs of PV systems after 2010.
Bio-energy, Biomass incinerators and fermentation
Mid-term Bio-energy Developments
SenterNovem (2005) follows market developments and bio-energy production as part of the Biomass
Action Plan. Table 9.11 shows the status of the bio-energy projects in 2004, as well as the actual
energy production in 2003. The figure also displays what the Biomass Action Plan expects the
contribution toward the different types of bio-energy production to be in 2010 based on preliminary
estimates by CBS and CertiQ. The Biomass Action Plan is a short-term action plan that is jointly
executed by the market and the government with the objective of improving the investment climate for
bio-energy so that more bio-energy projects can be realized.
Technology

Avioded fossil energy (TJ)

Large scale
Co firing
Waste incinerators
Small scale
Bio CHP incineration
Wood stove privates
Wood stove industrial
Landfill places
Bio CHP fermentation
Sewer waste treatment plant
Waste water treatment plant
Totaal

2001

2002

2003

2004
estimate

4555
12132

9960
12041

7123
12988

14000a
12703

1489
4842
2556
2309
84
2341
1083

1607
4842
2556
2022
150
2331
988

1550
4842
2556
1758
150
2129
898

31391

36497

33994

2280a,b
4842
2556
1560
400
2000
800
41141

Action plan
2010 (PJ)

34
20

8 - 18
7
2

4-6
75 - 87

Table 9.11: Contribution of bio-energy projects and Biomass Action Plan 2010, a estimates CBS and
CertiQ, b excluding cement ovens (Source: SenterNovem, 2005)

255

Part 5

The Biomass Action Plan shows that most of the developments will take place at large direct and
indirect co fired plants and small bio-CHP incinerators and are discussed hereafter. The planned
developments for the waste incinerators are discussed in the next section. Because there are no
significant developments planned for landfill dumps, biomass fermentation, and wastewater and
sewage treatment plants, the capacity of these types of energy generators is assumed constant in the
model calculations.
Large Direct and Indirect Co firing (Central)
The status of power stations with direct and indirect co firing is shown in Table 9.12.

Power Plant

Coal-fired
Maasvlakte
Amer
Borssele 12
Gelderland 13
Willem-Alexander
Hemweg 8
Gas-fired
Claus
Harculo
Eems
Total

Avoided
Avoided
fossil
fossil
energy
energy
2003 (TJ) 2002 (TJ)

3,966
1,389
1,216
452
38
26

3
34
7,124

Avoided Agreement
CO2 2003 2008 - 2012
(kton)
(kton CO2)

Use of
biomass
in 2003
(kton)

Approved
space
starting 2004
(kton)

4,578
3,081
1,334
184
25
-

373
131
114
43
3.6
2.5

805
931
310
466
200
488

234 81
76
33
4.9 1.6

759
-

0.2
1.9
-

n/a
n/a
n/a

0.1
0.9
-

1,600
0.463
-

9,961

669

> 3,200

432

3,020

1,200
120
60
40

Table 9.12: Status of power stations with direct and indirect co firing (Source: SenterNovem,2005)
Compared with 2002, avoided fossil energy decreased in 2003, among other things, mainly because
of the minimum financial support from the government as discussed earlier but also because the Amer
power station generated considerably less energy after its fatal scaffolding incident.
The current capacity (including Claus and Harculo) is not fully used but is sufficient to achieve the
goals in the agreement of coal between electricity production companies and government (see Section
2.2). About half of the agreed annual reduction of 6 Mtons CO2 can be met with the coal-fired power
plants alone. The expansion of biomass has a one-to-one relationship with the amount of MEP
compensation and the permits granted to generate electricity with use of direct and indirect co fired
bio-energy.
Small Bio-energy CHP Plants Based on Incineration (Decentralized)
At the end of 2004, about 18 small bio-energy plants (capacity < 50 MWe) were in operation, 6 of
which are incinerators and 12 of which are fermentation plants (SenterNovem, 2005). The
fermentation plants are usually small plants with a capacity of < 1 MWe. The incinerators often have a
higher capacity and play a modest role in the supply options. There are a number of initiatives to
create new incinerators, some of which have already been approved.
Among waste incinerator owners in particular, plans are being considered for the gasification or
incinerating of demolition wood on their grounds. The required incinerators are fairly large (15 - 25
MWe) and can contribute greatly toward the 2010 goals for decentralized bio-CHP incinerators.
According to SenterNovem, the current approved capacity (approximately 100 MWe) for small
initiatives is not sufficient to realize the goals in the Biomass Action Plan (approximately 250 MWe).

256

Chapter 9 Input Model for Developed Scenarios

Long-term Bio-energy Developments


The Projectgroep Transitie Biomassa (2003) submitted an elaborate vision on the role of biomass in
2040. Below is a summary of the most important conclusions that are related to this study. It is not
surprising that biomass was among the key power generation methods selected to spearhead the
transition. Biomass is mentioned in energy reports such as the government report Lange Termijn Visie
Energievoorziening (LVTE) (Long-term vision of energy supply), Shell's energy scenarios, and
Greenpeaces fossil-free energy scenario.
Biomass has wide applicability in a sustainable society (provided that production of biomass is (almost)
energy neutral). Large applications can be realized in different sectors, such as electricity and heat,
transportation, fuel, chemical, and materials.
The objective for 2040 is for bio-energy to cover 30% of the total Dutch energy consumption (600 to
1,000 PJ in a scenario in which saving energy is very important). Based on Europe's current policy on
bio-fuels and renewable electricity, the share of biomass in the Dutch energy supply will be some 3 to
4% in 2010. A subsequent annual growth of an average 7 to 8% above the increase in electricity
demand means that the level of 30% will be reached in 2040. Biomass will be used alongside other
renewable sources, such as sun, wind, and water.
In 2040, biomass will be supplied from a number of sources. In addition to using residual streams, the
Netherlands also cultivates biomass. Demand, however, exceeds the country's production capacity.
For this reason, biomass is procured on an international market. From a sustainability standpoint,
conditions are imposed on the cultivation of biomass in the areas of food supply, conservation,
biodiversity, and respect for local conditions in developing countries.
The groups that participated in the workshop Biomass Transition: On the road with a vision, developed
the different targets into transition paths. In order to apply some kind of structure, the transition paths
were clustered (see Figure 9.17.)

Transportation
fuels

Raw materials
Half manufactures

Electricity and
heat

Chemestry and
materials
Figure 9.17: Transition Biomass, 4 clusters (Source: Projectgroep Transitie Biomassa (2003))
The following transition paths are important for this thesis:
1) Pyrolysis
2) Biomass and coals
3) Syngas.

257

Part 5

Pyrolysis consists of converting biomass into bio-oil that can be used for indirect co-firing in coal-fired
and gas-fired power plants as well as in gasification plants, and for fuel for diesel engines and gas
turbines. Bio-oil can be produced where biomass is largely available and later transported to other
places in the world where it is converted into end products.
In 2040, indirectly co firing bio-oil in power stations is expected to avoid more than 10 Mton CO2
emission, an amount that could increase if bio-oil is used in the transportation fuel, chemical, and CHP
sectors.
Biomass and coal focus specifically on increasing the use of biomass to co-fire in coal-fired and gasfired power plants. In 2040, biomass is expected to contribute 200 PJ, coal 150 PJ, and gas a further
250 PJ. By then, 40% of the resources used to co-fire in coal-fired power plants will be biomass. In the
gas-fired power plants, this percentage will increase to 30%.
Syngas is based on a flexible and multifunctional intermediate product made of sources that contain
carbon, such as coal, natural gas, and biomass, such as wood from European forests and agriculture.
Wet and dry organic residual streams are also considered. Gasifying biomass creates bio-syngas, a
mixture of carbon monoxide (CO) and hydrogen (H2), from which a number of products can be made.
The advantage, however, is that they fit in the existing infrastructure. Bio-syngas has a considerable
potential to substitute primary energy that is contained in fossil fuels, see Table 9.13.
Sector
2010
2020
2040
Gas supply
90
220
980
13 (*2)
39 (*2)
204 (*2)
Transportation1)
Chemistry
18
39
117
Electricity sector
28
55
166
Total
150 - 160
350 - 400
1450 - 1650
Table 9.13: Potential of bio-syngas for various sectors 1) If syngas is also used for the production of
bio-ethanol, as a substitute for benzine, de demand for syngas doubles (Source: Projectgroep
Transitie Biomassa (2003))

Bio-energy, Waste Incinerators


Development of the Share of Supply over Time
The amount of electricity generated by waste incinerators has been stable for the period 2000 - 2005
mainly because the installed capacity has hardly increased.
Mid-term (2008-2012) / Long-term (2020-2025) Developments
According to the national waste management plans (LAP) policy scenario, the Netherlands will
produce an estimated 12 million tons of combustible waste in 2012 (LAP Part 3, Capacity plans). At
the moment, the incineration capacity is about 5.2 million tons. With 5.0 million tons of incinerated
waste in 2003 (CBS, 2007), the maximum incineration capacity has been reached. There is, however,
potential for expansion and a number of waste incinerators are considering extension of their
incineration capacity. Developments on the European waste market caused by and lack of
government support could upset these plans. Differences in legislation are preventing harmonization
on the international waste market and the past years this has put a stop to or postponed many waste
projects (SenterNovem, 2005). The tedious permit procedures and the absence of stimulation
measures (MEP scheme), are also an obstacle to these developments. However most of the waste
incinerators have expansion plans that are ready to be executed. Together, these plans represent a
total generation capacity of approximately 3.4 million tons and an increase of installed electrical power
of approximately 400 MWe. Only the expansions of the Amsterdam Waste and Energy Company
(AEB) and Huisvuilcentrale N-H in Alkmaar have been decided and are/have been build (situation
2007).

258

Chapter 9 Input Model for Developed Scenarios

9.3.5

Import/Export

General
The Netherlands relatively high amount of import capacity present-day (2008) consists of five crossborder connections: three with Germany and two with Belgium. The capacity has been expanded with
a new connection of 600 MWe to Norway and a project is developed for a new connection to England
(see also Part 2, Chapter 5). Because of the difference in electricity prices between the Netherlands
and the surrounding countries, a lot of electricity is currently being imported.
Factors that Influence the Share of Supply from Import/Export
The following factors influence the share of supply from import/export (see also under Economics):
- Available interconnector capacity;
- Difference in base and peak load prices between the surrounding countries;
- Cost of the interconnector capacity;
- Production capacity in surrounding countries.
Available Interconnector Capacity
The present-day interconnector capacity can already handle an import capacity of 5,000 MWe, but for
a number of reasons (see Section 5.2.5) the securely available capacity is between 3,350 and 3,850
MWe only. The further expansion of wind power in northern Germany and Germany's grid
reinforcement schedule mean that the securely available capacity will probably not be increased until
2007 (E.ON Netz, 2005). Between now and 2015, almost 850 additional kilometers of 380 kV
transmission routes are needed in Germany to transport wind power to the consumption centers in
central and southern Germany.
Once the NorNed cable, which can be expanded to 1,200 MWe, and the cable to England, which has
a capacity of 1,000 MW MWe, have been commissioned, the available interconnector capacity in off
peak periods will be as displayed in the below Figure 9.18. ECN (2005) states that the capacity
between the Netherlands and Germany and Belgium will be between 500 and 700 MWe lower in the
peak periods as a result of reliability criteria due to increased exchange on the European continent.

Figure 9.18: Development of off peak interconnector capacity (Source: ECN, 2005)

259

Part 5

Difference in Base and Peak Load Prices Between the Surrounding Countries
The import (or export) of electricity from surrounding countries is driven by the differences in base and
peak load prices. Figure 9.19 displays the price difference for base load and Figure 9.20 displays the
difference for peak load in the Netherlands, Germany and France.
65

60

55

NED BL 04

50
Base Load [/MWh]

NED BL 05
NED BL 06
45

GER BL 04
GER BL 05
GER BL 06

40

FRA BL04
FRA BL05
FRA BL 06

35

30

25

30
-1
220
05

30
-1
220
04

31
-1
220
03

31
-1
220
02

20

Figure 9.19: Base load electricity prices in the Netherlands, Germany and France (Source: Endex,
own information)
85

80

75

NED PL 04

70
Peak Load [/MWh]

NED PL 05
NED PL 06
65

GER PL 04
GER PL 05
GER PL 06

60

FRA PL04
FRA PL05
FRA PL 06

55

50

45

30
-1
220
05

30
-1
220
04

31
-1
220
03

31
-1
220
02

40

Figure 9.20: Peak load electricity prices in the Netherlands, Germany, and France (Source: Endex,
own information)

260

Chapter 9 Input Model for Developed Scenarios

Both figures clearly show that the German and French base and peak-load prices are lower then the
prices in the Netherlands but they follow the price developments in the Netherlands were the peak
load electricity prices are closely linked to the price of gas. The base load price differs less between
France, Germany, and the Netherlands than the peak load price. This means that the price difference
between the countries will be small during the off peak periods, which is obvious considering that the
price of electricity generated at coal-fired power plants as of 2000 till recent is determining the off peak
prices.
It is remarkable that in 2006 French electricity prices are higher than German prices. As a result,
Germany will increase its supply to France, which is the exact opposite of what happened in previous
years. The market prices in Belgium range between the Dutch and French prices.
Cost of Interconnector Capacity
The capacity on the high-voltage grid's five cross-border connectors is auctioned by TSO Auction, a
TenneT subsidiary. The available capacity is auctioned in daily, monthly, and annual auctions, during
which capacity is sold in both directions on each cross-border connector. The German TSOs, E.ONNetz and RWE Net, and the Belgian TSO, Elia, participate in the auction (TenneT, 2004). The division
of import capacity for auction was (situation per December 31, 2004) for the annual auction 900 MWe,
for the monthly auction 849 MWe, and the remaining capacity for the daily auction on the APX. If a
total of 3,350 MWe transborder capacity is available, and 900 MWe are reserved for long-term import
contracts (TenneT, 2006), 701 MWe was available for the daily auction. The costs for the annual and
monthly capacity for 2005 are specified below as an example.
From

Available Obtained
Price
(MW)
(MW)
(/MW)
328
328 18,921.60

To

ELIA

TenneT

TenneT

ELIA

RWE
Transportnetz TenneT
Strom

328

328

876.02

356

356 51,700.00

TenneT

RWE
Transportnetz
Strom

356

356

E.ON Netz

TenneT

216

216 51,868,01

TenneT

E.ON Netz

216

216

603.35

603.35

Table 9.14: Annual Auction 2005 (Source: www.tso-auction.org)


The profits of the auction are primarily generated from the import capacity required from Germany,
and to a lesser extent, Belgium. For example, if we look at what it costs to transport 1 MWe from RWE
Transportnetz Strom to TenneT, and if we assume that this capacity is fully required during peak hours
(4,064 hours), the price per MWh is as follows: 51,700.00 / (4,064 h x 1 MWe) = 12.7/MWh
Available
(MW)

Obtained
Price
(MW)
(/MW)

From

To

ELIA

TenneT

313

313 2,737.92

TenneT

ELIA

313

313

RWE

TenneT

377

377 2,901.44

TenneT

RWE

377

377

E.ON Netz

TenneT

159

159 3.087.63

TenneT

E.ON Netz

159

159

51.10
51.10
37.12

Table 9.15: Monthly Auction January 2005 (Source: www.tso-auction.org)

261

Part 5

The same exercise for January 2005, which assumes that this capacity is fully required in the peak
hours (352 hours), results in the following price:
2,737.92 / (352 h x 1 MWe) = 7.8/MWh
If we look at Figure 9.19, we can conclude that the costs of the annual and monthly capacity have
such a leveling effect that the advantage of price difference is almost cancelled out .
Supply of Generation Capacity in Surrounding Countries
For information about the supply of production capacity in the surrounding countries as of 2004, see
Van Eck et al (2004). Replacement investments are required in the surrounding countries, and the
overcapacity will continue to decrease. Moreover, Germany has to deal with the planned shutdown of
nuclear power plants based on its decision to abandon nuclear energy (E.on Netz, 2005) the same
applies for Belgium.
Development of the Share of Supply over Time
Figure 9.21 displays the development of import from 1995 onwards. For 2002 to 2005, the supply is
divided by source (see Figure 9.22). The large group of green electricity buyers combined with the
limited domestic production capacity of renewable energy means that a lot of renewable energy will be
imported (CBS, 2007).

20,000

18,000

16,000

Net Import Electricity [GWhe]

14,000

12,000

10,000

8,000

6,000

4,000

2,000

0
1995

1996

1997

1998

1999

2000

2001

2002

2003

Figure 9.21: Development of the net imported electricity (Source: CBS Statline)

262

2004

2005

Chapter 9 Input Model for Developed Scenarios

20,000

18,000

16,000

Net ImportElectricity [MWh]

14,000

12,000
Other
Biomass

10,000

Hydro
Wind

8,000

6,000

4,000

2,000

0
2002

2003

2004

2005

Figure 9.22: Net imported electricity and subdivision by source (Source: CBS, 2007)
Medium to Long-term Developments
The import capacity will further increase due to the planned expansion of cable connections between
the Netherlands and Norway, and between the Netherlands and England. The expansion of the
German 380 kV grids will increase the capacity for the existing cross-border connections.
It is not really possible to draw conclusions about the medium tot long-term development of production
capacity in the surrounding countries and the eventually import by the Netherlands. The plans to
phase out nuclear energy in Germany and Belgium, the replacement investments that must be made
for nuclear energy in France, the fact that England will move from a gas-exporting to a gas-importing
country, the water supply in Norwegian reservoirs, and all of the other energy market and regulating
issues, one of which is the absence of a level playing field (Van Eck et al, 2004), make this aspect
very uncertain.

263

Part 5

9.4

Filling in the Model for the Developed Scenarios

9.4.1

Introduction

For the defined future scenarios in Chapter 7, the data that serves as input for the model is given in
this section. With use of the story lines in the 4 described scenarios and the mid- and long term
developments for the supply side as described in the previous sections we have filled in the future
(2012 and 2025) installed capacities of the electricity production companies and other electricity
producers. The applied production options and the grow in electricity demand is thereby a function of
the applicable future scenario. In the next section, first the demand and supply side for the scenarios
is given, followed by the parameters applied for dispatch and fixed cost. Finally, the financial
parameters for the annual cost of existing -, annual cost (based on the new investments) of new build
and annual cost of replacement generation capacity are given. For each of the input variables, a
short description and motivation is given.
9.4.2

Demand and Supply Side

The input variables for demand and supply in the model based on the 4 described scenarios are given
in Table 9.16 for 2012 and table 9.17 for 2025.
Description

Unit

Scenario A

Scenario B

Scenario C

Scenario D

%
%

1.0
0.0

2.0
0.0

1.5
0.0

3.0
0.0

Electricity production companies


Conventional (incl peak )
Nuclear
IGCC
CCGT
CCGT CHP
CCGT DH

MWe
MWe
MWe
MWe
MWe
MWe
MWe

15,193
8,466
449
1,153
2,596
835
1,694

18,893
10,866
449
2,453
2,596
835
1,694

16,393
9,666
449
1,153
2,596
835
1,694

20,685
10,866
449
2,453
4,388
835
1,694

Other electricity producers


CCGT, GT+HRSG, B+ST CHP
CCGT DH
Gas Engines
Wind
Other Renewable
Waste Incinerators
Others

MWe
MWe
MWe
MWe
MWe
MWe
MWe
MWe

9,875
4,765
244
2,000
2,000
314
485
67

7,176
3,507
244
1,513
1,250
110
485
67

9,125
4,265
244
1,750
2,000
314
485
67

7,176
3,507
244
1,513
1,250
110
485
67

Import capacity
Germany, Belgium
Norway
UK

MWe
MWe
MWe
MWe

4,450
3,850
600
0

3,850
3,850
0
0

4,450
3,850
600
0

3,850
3,850
0
0

TJ
TJ
TJ
TJ

289,818
225,061
29,271
34,767

215,768
166,713
22,800
26,254

260,387
199,982
29,275
30,410

215,760
166,717
22,789
26,254

Electricity Demand
Yearly Growth
Shift Peak/Off Peak
Electricity Supply

Heat total
Proces Heat
District Heat
Greenhouse Heat

Table 9.16 Demand and supply input variables in the model for 2012

264

Chapter 9 Input Model for Developed Scenarios

Description

Unit

Scenario A

Scenario B

Scenario C

Scenario D

%
%

1.0
0.0

2.0
0.0

1.5
0.0

3.0
0.0

Electricity production companies


Conventional (incl peak )
Nuclear
IGCC
CCGT
CCGT CHP
CCGT DH

MWe
MWe
MWe
MWe
MWe
MWe
MWe

16,902
4,111
449
3,553
6,475
835
1,479

26,332
8,911
6,579
6,053
2,475
835
1,479

21,802
6,511
449
6,553
5,975
835
1,479

33,715
12,511
3,514
7,253
8,059
835
1,543

Other electricity producers


CCGT, GT+HRSG, B+ST CHP
CCGT DH
Gas Engines
Wind
Other Renewable
Waste Incinerators
Others

MWe
MWe
MWe
MWe
MWe
MWe
MWe
MWe

19,290
6,015
244
2,250
8,000
1,564
1,150
67

5,926
3,507
244
1,513
0
110
485
67

13,790
5,015
244
2,000
4,500
814
1,150
67

5,926
3,507
244
1,513
0
110
485
67

Import capacity
Germany, Belgium
Norway
UK

MWe
MWe
MWe
MWe

4,450
3,850
600
0

3,850
3,850
0
0

5,050
3,850
1,200
0

3,850
3,850
0
0

TJ
TJ
TJ
TJ

363,307
282,889
34,187
39,031

215,649
166,670
22,788
26,191

308,883
232,790
34,209
34,684

215,652
166,680
22,781
26,191

Electricity Demand
Yearly Growth
Shift Peak/Off Peak
Electricity Supply

Heat total
Proces Heat
District Heat
Greenhouse Heat

Table 9.17 Demand and supply input variables in the model for 2025
Electricity Demand
In all scenarios a growth in electricity consumption is assumed. This is based first on the expected
grow of the GDP as stated in Section 9.22 and Figure 9.1 and second on the fact that the share of
electricity consumption in the total energy demand will rise further in the future caused by
technological developments mainly on the household market but also in the industry.
Electricity Production Companies
The amount and sort of supply from electricity production companies is depending on the chosen
scenario and the amount of supply from other electricity producers which is required in the chosen
scenario. The use of advanced technologies like CO2 capture and storage for supply from electricity
production companies is depended on the scenario and the environmental issues which are applicable.
Economic considerations are of importance in the sort of chosen electricity supply options.
Other Electricity Producers
The amount and sort of supply from other electricity producers is depending on the chosen scenario. It
is obvious that a scenario with focus on environmental issues has more renewable and/or combined
heat and power supply options compared with a scenario with less or no environmental issues. The
use of advance technologies like storage of (wind) generated electricity is depended on the scenario
and the globalization (joined efforts) which are applicable.

265

Part 5

Import / Export
In all scenarios the import/export of conventionally generated electricity is set equal to zero. It is
unclear what developments for import/export will take place in the future for mid- and long term. This
depends on strategic behavior of the German electricity production companies what to do with the
remainder capacity (Tennet, 2000) but also on German governments decisions what to do with the
amount of nuclear capacity, replace it or not (see also Van Eck, 2007). For France the same question
on strategic behavior can be raised, but decrease of nuclear capacity is not likely to happen. In Tennet
latest monitoring report (Tennet, 2006) on security of supply in the period 2005 - 2013 it is stated that
the Netherlands will stay dependent on foreign production capacity in the coming period but that
dependency will decrease till 2010 caused by investments in new production capacity. Because of the
mothballing of existing units an increase of dependency is expected after 2013. In this thesis, however,
with the assumption of a perfect working electricity market (see Section 1.3.2) it is assumed that new
generation capacity, required to fulfill the future electricity demand, is built in the Netherlands. This
because of the geographic position of the Netherlands (close to the north sea (cooling water!)) and all
existing facilities for coal, gas and electricity transport (Van Eck et al, 2004). Only in the scenarios in
which environmental issues are of importance maximum hydro import from Norway during peak
periods is taken into account.
Heat
For process heat, the grow is dependent on the amount of extra installed CHP power which depends
on the chosen scenario. In case more CHP is installed, heat generated with heat boilers can be
replaced with heat generated with CHP. In Table 9.18 the possible grow amount is given and these
are taken into account depending on the chosen scenario of which 2012 and 2025 are modeled. If no
growth is applicable, the numbers of 2003 are used.

Electricity production companies


Other electricity producers
Total Heat

Heat Supplied in
2003 [TJ]

Heat Supplied in
2012 [TJ]

Heat Supplied in
2025 [TJ]

12,851
146,764
159,615

20,000
180,000 - 200,000
200,000 - 220,000

20,000
210,000 - 260,000
230,000 - 280,000

Table 9.18 Process heat demand at production in areas of electricity production companies and others
electricity producers
In Table 9.19 the possible grow amounts for district heat are given for the areas supplied from
electricity production companies and other electricity producers till 2025. Supply of heat is also
possible from refineries or other industrial processes for example in the Randstad Zuid area. Growth
amounts are taken into account depending on the chosen scenario of which 2012 and 2025 are
modeled and 2050 is described in appendix E. If no growth is applicable, numbers of 2003 are used.
City / Location

District heating areas


Randstad Noord
Randstad Zuid
Utrecht
KAN Gebied
Tilburg / Breda
Twente
Others
Total Heat

Heat Supplied in
2003 [TJ]

Heat Supplied in
2012 [TJ]

Heat Supplied in
2025 [TJ]

4,530
9,300
3,800
510
2,800
420
0
21,360

9,407
13,264
4,221
1,561
3,413
1,489
1,149
34,504

21,329
22,953
5,251
4,131
4,911
4,100
3,957
66,633

Table 9.19 District Heat demand at production in areas of electricity production companies and others
electricity producers (Source: Municipal Amsterdam, Rotterdam, Almere, ROM Rijnmond, own
information)

266

Chapter 9 Input Model for Developed Scenarios

Randstad Noord is the bundling of Amsterdam, Almere and Purmerend, Randstad Zuid is the bundling
of Rotterdam, The Hague and Leiden en KAN gebied is the area of Arnhem, Nijmegen. The difference
with the previous Table 9.1 is the generated heat from the waste incinerators, AVR in Duiven (KAN
gebied) and AEB in Amsterdam (Randstad Noord) and CCGT Marsteden in Enschede (Twente). The
extra amount of generated heat is 1,030 TJ on top of the in Table 9.1 mentioned 20,330 TJ..
For greenhouse heat, the grow is depended on the amount of extra installed small CHP power (gas
engines) which depends on the chosen scenario. For installed small CHP in greenhouses also
operating hours are of importance, the sharp decrease in operating hours as of 2004 means less heat
production. In this study the operating hours of 2003 are assumed to be applicable for the electricity
and heat production in 2012 and 2025. In case more small CHP is installed heat generated with heat
boilers can be replaced with heat generated with small CHP. In Table 9.20 the possible grow amount
is given and these are taken into account depending on the chosen scenario of which 2012 and 2025
are modeled. If no grow is applicable, the numbers of 2003 are used.
Heat Supplied in
Heat Supplied in
Heat Supplied in
2003 [TJ]
2012 [TJ]
2025 [TJ]
Electricity production companies1
Other electricity producers
Total Heat

0
26,322
26,322

0
30,000 - 35,000
30,000 - 35,000

0
35,000 - 40,000
35,000 - 40,000

Table 9.20 Greenhouse heat demand at production in areas of electricity production companies and
other electricity producers
9.4.3

Operation-, dispatch and fixed cost

Fuel prices
As can be noticed in Figure 9.19, 9.20, 9.24 and 9.25 price developments for electricity, gas and CO2
are unpredictable and they depend largely on the political and economical developments in the World.
The gas price is related with the oil prices and the electricity prices (present-day) with the gas prices
for peak and coal prices for off peak but they depend both also on CO2 prices, see Figure 9.25.
For the scenarios every price level for fuel and CO2 can be chosen. For this thesis we choose the
world oil prices reference case of the EIA (2007) in their Annual Energy Outlook (see Figure 9.23).

Figure 9.23 World Oil Prices in 3 cases 1990 2030 (Source : EIA, 2007)
1

For the electricity production companies no information was found on the amount of greenhouse heat delivered
by for example Roca 1, 2 and 3 and Amer 8 and Amer 9. Therefore the amount of heat generated by these
installations was included by district heating.

267

Part 5

In the model calculations the average forward price levels for 2006 in the second half of 2005 are used
for gas-, coal and CO2 prices to fix the prices differences between gas and coal (see Figure 9.24). In
combination with the reference case of the EIA the absolute price level for gas is determined for 2012
and 2025. It is assumed that forward gas prices for 2006, 2012 and 2025 are based on actual oil
prices in 2005, 2011 and 2024. As can be seen in Figure 9.23, the price levels for oil in the reference
case are more or less the same for 2005 i.e. $ 56.8/Barrel 2011 i.e. $ 54.3/Barrel and 2024 i.e. $ 55.6/
Barrel2. Therefore fuel price levels for all the scenarios are kept the same which is also an advantage
to compare the scenarios with each other.

90

80

80

70

70

60

60

50

50

40

40

30

30

20

20

10

10

Gasprice [ct/m 3], Coalprice [$/ton], CO 2 price [/ton]

90

3120
05
17
-1
-2
00
31
5
-1
-2
00
14
5
-2
-2
00
28
5
-2
-2
00
14
5
-3
-2
00
28
5
-3
-2
00
11
5
-4
-2
00
25
5
-4
-2
00
5
9520
05
23
-5
-2
00
5
6620
05
20
-6
-2
00
5
4720
05
18
-7
-2
00
5
1820
05
15
-8
-2
00
29
5
-8
-2
00
12
5
-9
-2
00
26
5
-9
-2
00
10
5
-1
020
24
05
-1
020
05
711
-2
0
21
05
-1
120
05
512
-2
00
5

Electricity price [/MWh]

For gas, the average price level second half of 2005 amounts approximately ct 22 /m3, see Figure
9.24, and with a lower heating value of 31.65 MJ/ m3, this leads to a price level of approximately 7
/GJ. In some other studies (ECN, 2005) a decrease of the gas price is foreseen to al level of 3.6/GJ
in 2012 - 2015. Because of the unpredictable price development as stated before and in combination
with the previous fuel scenario analysis the gas price level is assumed constant for the scenarios.

Peak 06
Base 06
Off Peak 06
Gas 06
Coal 06
CO2 06

Figure 9.24 Forward electricity, gas, coal and CO2 prices (Source: Endex, EnergieNed, own info.)
For coal average forward price levels for 2006 in the second half of 2005 where approximately
$ 60/ton, see Figure 9.24. With an average dollar/euro exchange rate of approximately 1.22 in the
second half of 2005 and a lower heating value of 24.8 GJ/ton, this leads to a price level of
approximately 2/GJ. Price levels for coal are assumed constant for the scenarios.
For nuclear fuel cost, briefing paper 8 of The Economics of Nuclear Power is applied. In this paper
the US $ cost to get 1 kg of uranium as UO2 reactor fuel is calculated. For 3,400 GJ thermal 315,000
kWh is generated with fuel cost of $c 0.52/kWh and this is with an average dollar/euro exchange rate
of 1.243 in 2006 approximately 1.2/GJ. Price levels for nuclear fuel are assumed constant for the
scenarios.
2

Present-day (2008) price levels of oil have increased sharply to levels above $ 100 per Barrel with even an
extreme of $ 147 per Barrel. Therefore also the gas price has increased significant (ct 22 /m3 to approx. ct 35
/m3 because these prices are related to the oil prices but also to dollar/euro exchange rate. Because the average
dollar/euro exchange rate increased also (1.2 to approx. 1.5) the gas price increase was relative less. Also the
coal prices went up sharply ($ 60/MT to above $ 150/MT). Prices difference between gas and coal, however, did
not really change. In case we would incorporated these price effects in the model, the absolute cost levels raise
but the relative differences do not change and therefore we did not change the input.

268

Chapter 9 Input Model for Developed Scenarios

For biomass, several fuel options are possible like, bio-oil, solid biomass, wood pellets, agro residues,
waste wood but also waste of other sectors for which incineration, in for example coal fired power
plants with high firing temperatures, is the most optimal removal (for example contaminated animal
meal, contaminated silt etc). For the last one, the burning has more social reasons than burning the
other biomass fuel options. Other options are mainly to generate green electricity, and therefore the
MEP schema is applicable. The cost of several biomass fuels (situation 2005) are given in Table 9.21.

Bio-oil gas plant (palm oil)


Fuel cost off Rotterdam
Additional transport
Energy content
Net fuel cost (range)

/ton
/ton
GJ/ton
/GJ

300 - 350
20
36,7
8,7 - 10,1 (9,4)

Bio-oil coal plant (palm oil)


Fuel cost off Rotterdam
Addition transport
Net fuel cost (range)

/ton
/ton
/GJ

250 - 350
20
7,4 - 10,1 (9,4)

Solid biomass gasification (coal plant)


Fuel cost net
Energy content
Net fuel cost (reference)

/ton
GJ/ton
/GJ

40
13
3,1

Wood pellets (coal plant)


Fuel cost net
Energy content
Net fuel cost (reference)

/ton
GJ/ton
/GJ

102
17,5
6,0

Agro-residues (coal plant)


Fuel cost net
Energy content
Net fuel cost (reference)

/ton
GJ/ton
/GJ

30 - 80
12,0
3,5

Waste wood (coal plant)


Fuel cost net
Energy content
Net fuel cost (reference)

/ton
GJ/ton
/GJ

20 - 70
14,0
3,5

Table 9.21 Cost of Biomass (Source: ECN, 2005)


As can be seen, the range of fuel cost is large where the cost of bio-oil which is imported is the most
expensive. The cost of bio oil is more than the cost of gas and the cost of bio oil and biomass is in all
cases more than the cost of coal. This means that without MEP scheme, bio oil and biomass is
economically considered not applied. For the scenario study it is assumed that in case stimulation
measures are applicable, bio oil is used as pure biomass in small scale biomass engines and
incinerators. Wood pellets are used as pure biomass in existing conventional and new conventional
coal fired multi fuel units. Solid biomass, waste wood, agro-residues (palm fibres, olive cake, cocoa
shells, soy bean etc.) and (not in the list) chicken dung are used as mixed biomass in the existing and
new IGCC multi fuel units.
For bio oil, a price level of 9.5/GJ is taken according to ECN (2005). For wood pellets, Kema (2005)
applies a price level range of 85 - 110/ton GJ with a typical price of around 100/ton. In combination
with a caloric value of 17 MJ/kg this leads to an average price level of 5.9/GJ. According to ECN the
price level for wood pellets is 6/GJ and this is also the price level used for the scenarios. Finally for
solid biomass, waste wood, agro-residues and (not in the list) chicken dung, mixed biomass, an

269

Part 5

average price level of 3.5/GJ is applied according to price levels mentioned in Table 9.21. For the
three types of biomass, price levels are assumed constant for the scenarios.
CO2 price
For CO2 the average price level second half of 2005 amounts approximately 22/ton, see Figure 9.25
and in 2006 the price level for 2007 is almost 20/ton again after a correction in April 2006. For the
scenarios a price level of 20/ton is applied and is assumed constant for the scenarios.

Figure 9.25 Correlation between price development CO2 en Endex contracts for baseload (Source:
Cozijnsen, 2006)
Startup cost
For the scenarios startup cost are only applicable for the generation units of the electricity production
companies as it serves as input for the PowrSym3 model (see Section 8.4).
Variable and fixed operation and maintenance cost
For the scenarios the variable operation and maintenance (O&M) cost are only applicable for the
generation units of the electricity production companies as it serves as input for the PowrSym3 model
(see Section 8.4).The fixed O&M cost are added in the main model and are calculated as function of
the generated electricity. For the generation units of the other electricity producers variable and fixed
O&M cost are combined and also calculated as a function of the generated electricity in the main
model. For the applied O&M cost see Section 6.5
Capacity and transport cost of natural gas
The capacity cost for the use of the main- and regional transport system by a generation unit is
depending on the maximum capacity required in relation with the total annual volume used by that unit.
De required contract capacity is the sum of the base load in which the base load capacity is defined as
annual volume divided by 8,760 hours. The remaining part is additional capacity (see Figure 9.26).

270

Chapter 9 Input Model for Developed Scenarios

Figure 9.26 Explanation cost of capacity and annual volume natural gas (Source: Cogen Projects,
2004)
The transportation cost for the use of the main- and regional transport system is depending on the
distance of the location of the gas fired generation unit form an entry point and the Groningen field.
Therefore, these cost are different for each location. Besides capacity and transportation cost, for the
use of the main- and regional transport system also a fee has to be paid in case the generation unit is
supplied via the regional distribution grid.
For the scenario calculations a tariff for capacity cost of 100 /m3/h is applied. For the generation
units of the electricity production companies, the cost for the main- and regional transport system is
calculated separately per unit and for the CHP units of the other electricity producers, these cost are
calculated for the cluster. The transportation cost are assumed to be 25% of the capacity cost (based
on own information). Cost for the regional distribution grid are only applicable for the (small scale)
CHP units of the other electricity producers because generation units of the electricity production
companies are commonly connected directly to the main or regional transport system. The cost for the
regional distribution grid are assumed to be 10% of the capacity and transportation cost (based on
own information).
9.4.4

Financial parameters

The financial parameters for the cost of existing-, new built- and replacement generation capacity are
required to calculate the annual financial cost. In combination with the dispatch and fixed cost we can
calculate the annual integral cost and the integral cost per MWh in order to compare the scenarios.
Annual financial cost of existing generation capacity
Electricity production companies
In Section 1.1 the switch of ownership of the Willem Alexander Power station and three of the four
electricity production companies is described. The valuation of the existing generation capacity
hereafter has been calculated with use of the take over prices as published during that time in annual
financial report and press statements. The EPZ stations were transferred to Essent and valuation of

271

Part 5

this capacity took place with use of investment levels of 2003 and thereby only applied for Amer 9,
CHP Moerdijk and Swentibold because of their commission data (1993, 1997 and 1999). For the other
stations of Essent the original investments are assumed to be depreciated. The nuclear facility in
Borssele is still owned by EPZ (50% Essent, 50% Delta) and the cost of the original investment is also
assumed to be depreciated.
For all the valuation a linear depreciation for a period of 15 year is assumed (EPZ, 2001; Essent 2002)
with an assumed WACC of 6%. In the calculations the depreciation starts one year after publication
date. The valuation levels are given in Table 9.22.
Publication

MNLG

MWe

/kWe

Take over Epon by Electrabel (80%), ING Bank (20%)


Take over Una by Reliant
Take over EZH by E.on
Take over Demkolec by Nuon
Essent / EPZ (Essent & Delta)

1999
1999
1999
2001

6000
4500
2350
225

2723
2042
1066
102

4,647
3,472
1,821
253
4,118
14,311

586
588
586
404

Take over Reliant by Nuon

2003

1088

3472

313

Valuation

Table 9.22 Valuation of existing generation capacity (Source : Electrabel, Nuon, EPZ, e.o. (see
reference list))
Other electricity producers
The valuation of the existing generation capacity of the other electricity producers has been calculated
with use of the investment levels of 2003, the commissioning date of the capacity and the applicable
10 year depreciation period. This means for industrial CHP, DH CCGT and small CHP that capacity
commissioned in 1993 and later has deprecation- plus capital cost (WACC of 6%) in 2003. Capacity
commissioned in 2002 and later has deprecation- plus capital cost in 2012 etc. The same is the case
for wind and waste incinerator capacity but depreciation periods are longer (15 and 20 years)
compared with industrial CHP, DH CCGT and small CHP capacity.
Annual financial cost of new build generation capacity
The annual financial cost of new build generation capacity for the scenarios has been calculated with
use of the investment levels of 2012 and 2025 (see Tables 6.7 to 6.12 of Section 6.5). The
commissioning date of new build generation capacity in the public utility situation was dictated by SEP
in order to maintain a safe reserve factor. Nowadays in the liberalized market the building of new
generation capacity is depending on other external influences (Van Eck, 2007) and commissioning
dates are unknown and so is the starting date of the depreciation in case of the linear depreciation
method. Therefore we used the annuity method, one factor for depreciation and interest. In Table 9.23
an overview is given of the applied depreciation periods for the several generation options, the annuity
factor at a WACC of 6% and the expected technical life time of the option.
Generation Options (new build)
CONVENTIONAL
IGCC
CCGT
NUCLEAR
CCGT, GT+HRSG, Small CHP
Wind, PV,
Waste Incinerator's

Periods

WACC

30
30
25
40
10
15
20

6%
6%
6%
6%
6%
6%
6%

Annuity Technical
0.0726
0.0726
0.0782
0.0665
0.1359
0.1030
0.0726

Table 9.23 Overview financial parameters for new build generation capacity (Source: Electrabel,
Essent, Nuon, EPZ, own information)
272

40
40
35
50
25
20
25

Chapter 9 Input Model for Developed Scenarios

Annual financial cost of replacement generation capacity


As can be noted in Table 9.23, the technical life time of the generation options can be extended
towards even 50 years for the nuclear option. Of course this means that investments have to been
done for lifetime extensions programs and this will have influence on the annual financial cost. For
this thesis however assumptions had to be made and these are summarized bellow:
For Conventional, IGCC, CCGT units, the cost of modification packages for boiler, steam
turbine and gas turbines, HRSG, steam turbine are assumed to be included in the O&M cost
unless other (investment) information is available (see Table 9.4 modernizations plans);
For the Nuclear unit in Borssele the cost of the modification package was not taken into
account;
For industrial CCGT and GT+HRSG the cost of modification packages for gas turbines,
HRSGs, steam turbines are assumed to be included in the O&M cost;
For small CHP (gas engines) it is assumed that the gas engines are replaced after 15 years of
operation by a completely modified or new gas engines at a rate of 50% / 50%. The cost of
new gas engines amounts to half of the original investment cost. For this thesis the cost of
modified and new gas engines are assumed to be included in the O&M cost.
For wind, PV no modification packages are foreseen;
For waste incinerators modification packages for ovens, boilers, steam turbines are assumed
to be included in the O&M.

9.4.5

Stimulation measures

The cost of the stimulation measures for the renewable options and CHP in 2012 and 2025 is
calculated with use of the MEP rebate for the second half of 2006 as given in Table 9.3. This is
because the MEP rebates have to be calculated with use of the forward fuel-, electricity and CO2
prices in 2005 for 2006. The rebate for on- and offshore wind energy is corrected for the lower future
investment levels and operation and maintenance cost. The rebate for installations > 50 MWe, the
central co-firing, is corrected for the (variable) rebate (if required) to decrease the cost of biomass in
/GJ to a price level required to achieve a maximum amount of operating hours and produced
electricity of which 20 and 40% is generated with biomass as fuel.

9.4.6

Parameters not taken into account

Fiscal rules
Besides the stimulation measures also a few fiscal rules are applicable which can be positively applied
mainly for projects with CHP (Cogen Projects, 2004). The most important fiscal rules are:
EIA, energy investment allowance, applicable for CHP units with a Senter efficiency > 65%,
gives a direct financial fiscal advantage because a percentage of 44% in 2005 of the
investment (Cogen Projects, 2004) can be deducted of the assessable income so that less tax
has to be paid;
VAMIL, earlier depreciation environmental investments, applicable for environmental friendly
greenhouses in which small scale CHP can be applied, earlier depreciation in case a company
creates revenues less tax has to be paid;
Green label greenhouse and green financing, applicable for environmental friendly projects,
for example greenhouses in which small scale CHP are applied. With a green label, a
greenhouse or plant owner can be considered for green financing for which interest rates for
rented capital are 1 to 2% lower than the market rates.
These fiscal rules are project dependent and its continuity for the period 2012 - 2025 is unpredictable.
In case of EIA, not all new CHP projects can make use of the advantages of the fiscal regulations
because a Senter efficiency (e + 2/3th ) > 65% cannot be reached. The VAMIL and green labels are
mainly applicable for the projects with small scale CHP. The influence on the total annual system cost
are assumed to be very low and therefore fiscal rules are not taken into account for the scenarios.

273

Part 5

Grid connection and backup cost of electricity


Each unit is coupled with the electrical grid, for supply of electricity to the grid but also to use electricity
of the grid in case the installation is not in operation (see section 5.2). The larger units of the electricity
production companies are mostly direct connected to the 380 and 220 kV transmission grids of the
TSO. The units of the other electricity producers are mostly coupled to the distribution grid of the
regional grid company (150 kV and lower3). The cost of grid connection and backup electricity can be
split-up:
Grid connection
o Transport- and connection fee;
o Metering cost;
Backup electricity
o Commodity electricity cost;
o Energy tax.
The cost for transport of electricity over the grid is depending on the voltage level the installation is
connected to the grid. These cost are split-up in fixed fee, contracted capacity (kWe), maximum
capacity during a month, cost per used kWh in peak and off peak. Besides these costs also system
service costs to balance the national grid are applicable. In case the unit is connected to the
distribution grid, the regional grid company collects these service costs for the TSO.
The costs for metering are costs for the installation of the meter, maintenance (for example calibration)
and the (automatic) readout of the measured data. In case the unit is connected to the distribution grid
the regional grid company normally collects these costs for the certified metering company.
Electricity of the grid delivered by an energy company is required for own consumption (heat,
ventilation, air condition etc.) in case the unit is out of operation. For this electricity a staffed energy tax
is applicable. In most units the electricity from the grid is also used to startup the installation with use
of its own generator. For CHP units delivering to an industry, electricity has to be delivered in case the
CHP is planned or unplanned out of operation and a heat boiler is producing the heat. Depending on
the required capacity (kWe) and time of outage of the CHP this can be a large amount of electricity.
The cost of the grid connection is different for each unit connected to the electrical grid and it is not
easily possible to calculate the total annual cost system cost. The influence however on the total
annual system cost is assumed to be very low and therefore costs of grid connection are not taken
into account for the scenarios.
As stated above, backup costs of electricity are mainly applicable in case of a planned or unplanned
shutdown of a CHP unit. Costs however can vary (ECN, 2005) depending of the duration of the stop.
Short stops can be relative expensive, specially in case the stop is unplanned; for the longer stops,
cheaper electricity purchase solutions can be found. Because this thesis is focusing on the costs of
the total production system and these costs are related with the consumer side, backup costs
electricity are not taken into account.
Energy tax natural gas
For natural gas used in CHP installations an exemption energy tax natural gas is applicable in case
the electrical efficiency of the installation is at least 30%. For almost all CHP installations the electrical
efficiency is higher then 30%. Standalone heat boilers do not have this exemption. The energy tax
tariff is a so called staffed tariff per m3 natural gas, the tariff decreases in case more natural gas is
used. The influence of these energy tax on the total annual system cost therefore are assumed to be
very low and therefore the they are not taken into account for the scenarios.
Cost of infrastructure for District Heating
For District Heating, a more expensive infrastructure is required compared with an infrastructure for
the supply of gas. Also more investments have to be done in extra DH related equipment (heat
exchangers, pumps, heat buffers, piping etc.) in generation units for producing the heat (Van Eck,
2007). The consumers using district heating however pay on average not more than if they should
3

As of January 1, 2008 the management and control of the 110 and 150 kV grids is transferred from the regional
grid companies to the TSO as result of the new Law Onafhankelijk netbeheer (Independent grid management)
(Min EA, 2007).

274

Chapter 9 Input Model for Developed Scenarios

have had a gas connection and a heat boiler. This so called niet-meer-dan-anders (NMDA) principle
(not more then else), is formerly established by the government in the tarief advies (tariff advice) for
district heating delivery (EnergieNed, 2005). The momentary applied NMDA exists of (Algemene
Rekenkamer, 2007):
1. Earning capacity fee (once);
2. Connection fee (once);
3. Fixed fee (yearly)
4. Heat price (yearly), depending on used heat.
The earning capacity fee is applied once and is a competition and environmental based fee. Because
with district heating a better environmental performance can be reached compared with a individual
heat boiler installation, owners can avoid energy reduction measures and cost in their houses to fulfill
the EPN norm and this justifies, according to the tarief advies the earning connection fee.
The connection fee has to be paid the moment a house is connected to the district heat grid. These
costs are based on avoided costs of a gas connection and of the difference between investment of a
heat boiler installation and a district heat installation.
The fixed fee has to be paid yearly for the avoided fixed fee of a gas connection, avoided costs of
maintenance and spare parts of a heat boiler installation and costs for difference in technical life time.
The heat price is calculated based on the marktwaarde principe (market value principle). The heat
price based on this principle means for the customer that his yearly cost for delivered heat is never
higher than the average cost in case gas was used.
Because three of the four components are based on the avoided gas situation, only the earning
capacity fee can be related with the more expensive infrastructure for district heating for the scenarios.
For new project energy companies, municipals, project developers negotiate about the values of this
fee. The earning capacity fee is the variable influenced by the negotiations about the tariffs. The use of
this fee for the new district heat connections in the scenarios as extra costs for district heating
connections is however not easily possible. The fee is project and energy company dependent and
information about the value of this one in the acquisition of a house discounted fee, are therefore not
public and difficult to get (Algemene Rekenkamer, 2007).
The influence of the more expensive infrastructure for district heating on the total annual system cost
is assumed to be low. The depreciation time for these investments is long (30 - 40 years) and in case
a 40 year depreciation is applied with a WACC of 6%, this means an annuity of 0.06. Therefore the
extra cost compared with heat boiler systems and extra cost for the heat production for District
Heating are low and not taken into account for the scenarios.
9.5

Summary

Data that serves as input for the scenario calculations with the model are specified. The Dutch
electricity demand is analyzed and an electricity load curve is derived. The development of the share
in total generation of the individual modules over time and the expected mid-term (2008 - 2012) and
long-term developments (to 2020 - 2025) are analyzed and input for the supply side for each scenario
is derived. Finally, the financial parameters applied for the economic calculations are described and
explained.

275

Part 5

276

Chapter 10 Analysis of Scenario Outcomes

Chapter 10
10

Analysis of Scenario Outcomes


10.1

Introduction

The main objective of this research was to ascertain whether it is possible to strike an optimal balance
between environment, economics and security of supply in the future energy situation. The question
that needed to be answered was whether an optimal balance could be achieved in the perfectly
functioning electricity market, which was assumed for the purposes of this research, or whether steps
need to be taken by the government and the business community. Parts 1 & 2 of this thesis describes
the existing energy supply system. Part 3 looks at future developments that will influence the supply
side in particular. Part 4, Chapter 7 presents four possible future scenarios which cover trends in
supply and demand, the question of protectionism versus globalization, and the importance of the
environment. The four scenarios are simulated with the techno-economic simulation model described
in Part 4, Chapter 8 for which the input parameters were defined in Part 5, Chapter 9. In Chapter 10
the results of the simulations are analyzed to ascertain the balance between environment, economics
and security of supply for each scenario. The results are then compared and these form the starting
point for the conclusions and recommendations in Chapter 11.
10.2

Model Results for the Developed Scenarios

Figures 10.1 - 10.6 present part of the calculated results for the defined scenarios. Extensive and
detailed reporting on the outcomes are given in appendix F.
10.2.1 Environment
140,000

120,000

CO2 [kTon]

100,000

80,000
Scenario A
Scenario B
60,000

Scenario C
Scenario D

40,000

20,000

0
2000

2005

2010

2015

2020

2025

2030

Figure 10.1 Total CO2 emissions

277

Part 5

800

700

600

CO2 [g/kWh]

500

Scenario A

400

Scenario B
Scenario C
Scenario D

300

200

100

0
2000

2005

2010

2015

2020

2025

2030

Figure 10.2 CO2 emission per MWh

10.2.2 Economics
11,000
10,000

Dispatch + Fixed + Financial Cost [M/a]

9,000

8,000
7,000

6,000

Scenario A
Scenario B

5,000

Scenario C
Scenario D

4,000
3,000

2,000
1,000
0
2000

2005

2010

Figure 10.3 Total annual integral cost

278

2015

2020

2025

2030

Chapter 10 Analysis of Scenario Outcomes

70

Dispatch + Fixed + Financial Cost [/MWh]

60

50

40

Scenario A
Scenario B
Scenario C
Scenario D

30

20

10

0
2000

2005

2010

2015

2020

2025

2030

Figure 10.4 Specific integral cost

30000

Cummulative investment cost [M]

25000

20000

Scenario A

15000

Scenario B
Scenario C
Scenario D

10000

5000

0
2000

2005

2010

2015

2020

2025

2030

Figure 10.5 Total capital expenditure

279

Part 5

10.2.3 Security of Supply


14,000

12,000

Capacity [MWe]

10,000

8,000
Scenario A
Scenario B
Scenario C

6,000

Scenario D

4,000

2,000

0
2000

2005

2010

2015

2020

2025

2030

Figure 10.6 Available capacity vs. max load


10.3

Analysis of the Results for 2025

10.3.1 Scenario Results


Scenario A
The focus in scenario A (Section 7.3.4) is on the development of renewable energy sources, a
significant decrease in CO2 emissions, the transition to an entirely CO2 emissions free energy supply
around 2050, and handling the national gas reserves with care to ensure sufficient natural gas
availability for domestic consumption in 2050.
Environment
The government decision to stimulate renewable energy sources leads to 6,000 MWe of installed wind
power capacity at sea, 2,000 MWe on land, and 1,500 MWe of photovoltaic capacity. The share of
renewable energy sources is almost 27% of total installed capacity.
The use of domestically produced biomass, including organic waste, increases to more than 10% of
the total required fuel consumption for electricity and heat supply. This is achieved mainly by
deploying large amounts of domestically produced biomass fuel in conventional and IGCC power
plants.
CCS systems are only applied to a modest extent in one conventional and two IGCC units.
The industry is stimulated to generate more process heat using CHP, which results in an increase of
CHP capacity by 3,000 MWe up 40% compared with the reference situation in 2003. The additional
CHP and industrial heat capacity for process- and district heat helps to avoid approximately 5 billion
m3 of gas consumption and 9 Mtons of CO2 emissions by conventional boiler systems.
District heating is implemented on a large scale both in existing and new areas and leads to a 150%
increase in connections for houses and small and medium-sized enterprises compared with the
reference situation in 2003. Here, the use of waste heat of power plants and industries is stimulated as
much as possible.

280

Chapter 10 Analysis of Scenario Outcomes

Due to these measures the total CO2 emissions of the supply system drops to 48 Mtons which is a
reduction of 35% compared with the reference situation in 2003. The emission level for each MWh
produced is 350 g/kWh.
Economics
Due to the relatively high costs of installing CO2-emission-poor generating capacity, capital
expenditure in this scenario are high, both in absolute and in relative terms, compared with the other
scenarios although electricity demand increases by just 1% a year. This is caused mainly by the
increase in installed renewable energy sources which show higher investment levels, in spite of
learning curve effect, compared with conventional power supply sources.
In the given market situation, the 8,000 MWe of wind power can only contribute to the total electricity
supply if a significant government subsidy is made available. In far less extent the same holds for the
application of biomass fuel in conventional and IGCC power plants for extra investment cost in the
installation. The total integral cost price of electricity in this scenario to amounts 60 /MWh with a total
annual subsidy budget of 1.6 billion/a.
Security of Supply
In this scenario, 8,000 MWe of wind power and 1,500 MW of photovoltaic capacity make a significant
contribution to total electricity supply however, without large electricity storage systems. Reserve
capacity (reserve factor of 36%) is required to compensate for the wind variability between 0 and
8,000 MWe and to avoid a shortfall in electricity supply on windless days in combination with a high
electricity demand.
Large-scale CCGT power plants have been built because of their excellent controllability between part
and full load operation and their short startup times. However, operation times of the CCGT power
plants are short and therefore cannot be used to supply process and/or district heat.
Occasionally, more wind power is supplied than is required by the system. The combination of a large
supply of wind power and low demand for electricity leads to oversupply of wind power, causing the
wind turbines to run in part load even when all power plants are running in part load or are switched off
where possible. The wind energy loss however is low, less than 0.2% of the total electricity demand in
a year.
The use of natural gas for electricity and heat supply increases to 22.1 billion m3/a versus 17.6 billion
m3/a in 2003 due to the higher demand for electricity, the larger share of CHP in the supply system,
and the gas consumption of the large-scale CCGT power plants. Domestically, the gas consumption
by individual boilers for heat supply decreases by approximately 5.0 billion m3. As a results total
combined gas consumption will not change compared to 2003.
Scenario A summarized:
Environment - favorable in terms of CO2 emissions
Economics - high integral electricity cost price
Security of supply - only safeguarded after the installation of significant additional CCGT capacity.
Scenario B
The focus in this scenario (Section 7.3.4) is on electricity generation at minimum cost, no new
developments in renewable energy, and investments in CHP only where a fast return on investment is
guaranteed. The national gas reserves are used in pursuit of gains and availability for domestic
consumption in 2050 is just enough to cover the domestic gas needs for indoor heating and hot tap
water in areas without district heating.
Environment
The government decision to no longer stimulate renewable energy sources forces utilities to
discontinue their investments in new wind power capacity. Existing wind turbines are not replaced and,
as a consequence, wind capacity at sea and on land drops to 0 MWe. The share of renewable energy
sources in the form of existing photovoltaic capacity and hydro power amounts to no more than 0.4%
of total installed capacity and the use of biomass, including organic waste, decreases compared with
2003.

281

Part 5

The share of nuclear-generated electricity increases to over 30% of total electricity production and that
of coal-generated electricity even to 43% but without introducing CCS system.
As the industry was not stimulated to generate more process heat using CHP, CHP capacity stayed
the same as in 2003 and only replacement investments were made. District heating continues only for
existing projects with the focus on cost reductions.
Total CO2 emissions of the supply system increases to 81 Mton, up 10% compared with the reference
situation in 2003. Emission level for each MWh produced is 471 g/kWh.
Economics
Capital expenditure in this scenario is by far the lowest compared with the other scenarios although
electricity demand increases by 2% each year. This is caused mainly by the absence of newly
installed renewable energy capacity, no increase in CHP capacity, and the large-scale use of
conventional and IGCC power plants without CCS installations and nuclear power plants.
Total integral cost price of this scenario amounts to 47 /MWh. There is no government subsidy in this
scenario.
Security of supply
In this scenario without large-scale deployment of wind power and photovoltaic capacity, a reserve
factor of 10% is required to compensate for planned and unplanned loss of 3 large generation plants
and to avoid a shortage in electricity supply in case of high electricity demand.
The use of natural gas for electricity and heat supply decreases to 14.5 billion m3/a versus 17.6 billion
m3/a in 2003 due to the larger share of nuclear and coal generated electricity in the supply system.
Domestically, gas use by individual boilers for heat supply remains flat and, as a consequence, total
combined gas consumption decreases with 3.3 billion m3/a compared to the reference situation in
2003.
Scenario B summarized:
Environment - less favorable in terms of CO2 emissions
Economics - low integral electricity cost price
Security of supply - safeguarded with a modest amount of additional capacity.
Scenario C
The focus in scenario C (Section 7.3.4) is on the development of renewable energy sources, a
significant decrease in CO2 emissions, the transition to an entirely CO2 emissions free energy supply
in 2050 in a society with far-reaching international collaboration and economic globalization. The
national gas reserves are handled with care to ensure sufficient natural gas availability for domestic
consumption in 2050.
Environment
The government decision to stimulate renewable energy sources leads to 3,000 MWe of installed wind
power capacity at sea, 1,500 MWe on land, and 750 MWe of photovoltaic capacity. The share of
renewable energy sources is almost 15% of total installed capacity.
The use of domestically and internationally produced biomass, including organic waste, increases to
over 10% of the total required fuel consumption for electricity and heat supply. This is achieved mainly
by using large amounts of biomass fuel in conventional and IGCC power plants.
CCS systems are being deployed in all of the conventional and IGCC units built between 2012 and
2025.
The industry was stimulated to generate more process heat using CHP and this results in an increase
in CHP capacity by almost 1,800 MWe, up 23% compared with the reference situation in 2003. The
additional CHP and industrial heat capacity for process - and district heat helps to reduce gas
consumption by approximately 3.2 billion m3 and CO2 emissions of conventional boiler systems by
5.7 Mtons.
District heating is implemented on a large scale both in existing and new areas leading to a 150%
increase in connections of houses and small and medium-sized enterprises compared with the
reference situation in 2003. Here, the use of waste heat of power plants is stimulated as much as
possible.

282

Chapter 10 Analysis of Scenario Outcomes

Due to these measures total CO2 emissions by the supply system drops to 51 Mtons, down almost
31% versus 2003. Emission levels for each MWh produced is 329 g/kWh.
Economics
Due to the relatively high cost of installing CO2 emission poor generating capacity, the capital
expenditure of this scenario are high compared with the other scenarios although electricity demand
increases by only 1.5% a year. This is caused mainly by the increase in installed renewable energy
sources which show higher investment levels, in spite of learning curve effect, compared with
conventional power supply sources. Also, the large number of installed CCS installations where
learning curves are expected to have impact on investment levels after 2025 have an increasing effect
on capital expenditures.
In the given market situation, the 4,500 MWe of wind power can only contribute to total electricity
supply if a significant government subsidy is made available. In far less extent the same holds for the
application of biomass fuel in conventional and IGCC power plants for extra investment cost in the
installation. The total integral cost price of this scenario amounts to 56 /MWh with a total annual
subsidy budget of 0.8 billion/a.
Security of supply
In this scenario, 4,500 MWe of wind power and 750 MW of photovoltaic capacity will make a
significant contribution to total electricity supply however, without large electricity storage systems.
Reserve capacity (reserve factor of 17%) is required to compensate for the wind variability between 0
and 4,500 MWe and to avoid a shortfall in electricity on windless days in combination with high
electricity demand.
Large-scale CCGT power plants have been built because of their excellent controllability between part
and full load operation and their short startup times. However, operation times of the CCGT power
plants are short and therefore cannot be used to supply process and/or district heat.
In this scenario, as opposed to scenario A, the supply of wind power will not be higher than the system
requires. This is due to the higher electricity demand and the lower amount of installed wind capacity
compared with case A.
The use of natural gas for electricity and heat supply increases to 18.8 billion m3/a versus 17.6 billion
m3/a in 2003 due to the higher electricity demand, the larger share of CHP in the supply system, and
gas consumption by the large-scale CCGT power plants. Domestically gas use by individual boilers for
heat supply decreases by 3.2 billion m3. As a consequence total combined gas consumption
decreases with 1.9 billion m3/a compared to the reference situation in 2003.
Scenario C summarized:
Environment favorable in terms of CO2 emissions
Economics high integral electricity cost price
Security of supply only safeguarded after the installation of additional CCGT capacity.
Scenario D
The focus in this scenario (Section 7.3.4) is on electricity generation at minimum cost, no new
developments in renewable energy, and capital expenditures on CHP only where a fast return on
investment is guaranteed in a society with far-reaching international collaboration and economic
globalization.
Environment
The government decision to no longer stimulate renewable energy sources forces utilities to
discontinue their investments in new wind power capacity. Existing wind mills are not replaced and, as
a consequence, wind capacity at sea and on land drops to 0 MWe. The share of renewable energy
sources in the form of existing photovoltaic capacity and hydro power amounts to no more than 0.3%
of total installed capacity and the use of biomass, including organic waste, has decreased compared
with 2003.
Electricity generated with conventional coal and IGCC power plants, without introducing CCS systems,
increases to over 60% and the share of nuclear generated electricity to 13% of total electricity
production.

283

Part 5

As the industry is not stimulated to generate more process heat using CHP, CHP capacity stays the
same as in 2003, and only replacement investments are made. District heating is continued only for
existing projects with the focus on cost reductions.
Total CO2 emissions of the supply system increases to 120 Mtons, up 63% compared with the
reference situation in 2003. Emission levels for each MWh produced are 565 g/kWh.
Economics
Total capital expenditures in this scenario is highest compared with the other scenarios mainly due to
increase in electricity demand by 3% a year and the required of new installed capacity. There is no
newly installed renewable energy capacity, no increase in CHP capacity but only large-scale use of
conventional, IGCC, CCGT and nuclear power plants.
The total integral cost price of this system amounts to 49 /MWh. There is no government subsidy in
this scenario.
Security of supply
In this scenario, without large-scale deployment of wind power and photovoltaic capacity a reserve
factor of 10% is required as in scenario B to compensate for planned and unplanned loss of 3 large
generation plants and to avoid a shortfall in electricity supply in case of high electricity demand.
The use of natural gas for electricity and heat supply decreases to 16.5 billion m3/a versus 17.6 billion
m3/a in 2003 due to the larger share of nuclear and coal generated electricity in the supply system.
Domestically, gas use by individual boilers for heat supply remains flat and, as a consequence, total
combined gas consumption decreases with 1.3 billion m3/a compared to the reference situation in
2003.
Scenario D summarized:
Environment unfavorable in terms of CO2 emissions
Economics low integral electricity cost price
Security of supply safeguarded with a modest amount of additional capacity.
10.3.2 Scenario Results Compared
In addition on the Figures 10.1 - 10.6, the scenario results are compared on the differences between
the balance in Table 10.1 - 10.3 and Figure 10.7 - 10.9.
Environment
Security of supply
Scenario
Economics
Green electricity
Capacity - Max.Load
Integral Cost
[%]
[MWh]
[/MWh]
Scenario A
Scenario B
Scenario C
Scenario D

27
1
18
1

60
47
56
49

12,923
3,377
9,638
3,900

Table 10.1 Scenario results compared on absolute results (1)

Scenario

Environment
CO2 emissions
[Mton/a]

Economics
Investment cumm
[M]

Security of supply
Reserve factor (1)
[%]

Scenario A
Scenario B
Scenario C
Scenario D

47.8
80.6
50.6
119.8

20,833
17,809
21,653
19,644

36
10
27
10

Table 10.2 Scenario results compared on specific results (2)

284

Chapter 10 Analysis of Scenario Outcomes

Scenario

Environment
CO2 emissions red.
[%]

Economics
Stimulation
[M/a]

Security of supply
Reserve factor (2)
[%]

Scenario A
Scenario B
Scenario C
Scenario D

35
-10
31
-63

1,570
0
811
0

19
10
17
10

Table 10.3 Scenario results compared on specific results (3)

Green Electricity (%)


60
50
40
30
20
Reference 2003
Scenario A

10

Scenario B
Scenario C

Reservefactor (%)

Scenario D

Integral Cost [/MWh]

Figure 10.7 Scenario results compared on absolute results


CO2 emissions (% of max)
100%
90%
80%
70%
60%
50%
40%
30%
20%

Reference 2003

10%

Scenario A

0%

Scenario B
Scenario C
Scenario D

Reservefactor (% of max)

Investment (% of max)

Figure 10.8 Scenario results compared on relative results

285

Part 5

Finance
O&M
Startup, Gas Capacity, CO2 Transport
60

CO2
Operation
17

50

15

Cost components [/MWh]

10
11

40

12
3
30

7
2

1
6

8
9

20

3
0

10

19

23

19

23

11

20

0
Reference

Scenario A

Scenario B

Scenario C

Scenario D

Figure 10.9 Breakdown of specific integral cost in 2025


The impact of the supply choices in the various scenarios on the different cost components of the
integral cost is clearly shown in Figure 10.9. As might be expected, financing cost and cost of
operation and maintenance per MWh for Scenarios A and C are highest due to the large share of CO2emission-poor generating capacity. The cost of CO2 are lower for Scenarios A and C because of the
large share of renewable energy and gas fired supply. The specific operating cost for A and C are
higher due to the high volume of gas used as fuel in the decentralized power plants, the extra fuel
needed for CCGT power plants to compensate for the variations in supplied wind energy, and the
lower amount of generated electricity versus Scenarios B and D.
10.4 Effects of Electricity Storage for Scenario C
Electricity storage is one of the options to compensate for a surplus or shortage of wind energy (see
Section 6.4.5). Using the developed simulation model electricity storage based on a pumped storage
hydropower system (PAC) can be simulated. For Scenario C, Figures 10.10 and 10.11 illustrate the
effect of deploying a PAC with a storage capacity of 30 GWh and a load/extract capacity of 1,500
MWe on the normal load profile.

286

Chapter 10 Analysis of Scenario Outcomes

1,600

30

Pump
Generation

1,400

Energy PAC

25

1,200

15

800

Level PAC (GWh)

Generation/Pump (MWe)

20
1,000

600
10
400
5
200

0
1

15

22

29

36

43

50

57

64

71

78

85

92

99

106 113

120 127

134 141

148 155

162

Hour [h]

Figure 10.10 Generation / Pump hours and Level PAC moments week-1 January
30

18,000
No PAC

16,000

PAC
Wind
Level PAC

14,000

20
Level PAC [GWh]

Load [MWe]

12,000

10,000

8,000

6,000

10

4,000

2,000

0
0

24

48

72

96

120

144

0
168

Time [h]

Figure 10.11 Load central electricity production units with and without PAC week-1 January

Figure 10.11 clearly shows the effect of deploying a PAC facility on the load of the central electricity
production. The peaks in the load are flattened and the sinks in the load are filled up. The effects of a
sudden decrease in wind supply are visible between 110 - 130 hours. A drop in wind supply, normally
offset by electricity supply by CCGT installations, is compensated by extra electricity generation with
PAC and this explains the decrease in the PAC energy level.
287

Part 5

The energy balances of the electricity and heat supply systems with and without PAC are presented in
Table 10.4. The outcomes show that the presence of a PAC increases electricity demand due to a
double efficiency effect, first as a result of pumping (storage) and second due to generation. The
additional demand for electricity is completely generated with conventional and IGCC units decreasing
the operating hours of CCGT power plants. Using the simulation model these and other effects of
deploying a PAC on the Dutch electricity supply system can be studied in more detail but this is
beyond the scope of this thesis.
Electricity
PAC

NO PAC

Fuel
DIFFERENCE

PAC

NO PAC

CO2
DIFFERENCE

PAC

NO PAC

DIFFERENCE

Conv Coal
Nuclear
IGCC Coal
CCGT

30,001,919
3,782,184
44,377,193
15,491,058

MWh
29,619,730
3,780,037
43,404,329
16,231,817

382,189
2,147
972,864
-740,758

270,850
42,024
365,966
133,033

TJ
267,628
42,001
359,149
137,427

3,222
23
6,816
-4,393

18,423
0
9,530
8,705

kton
18,533
0
9,330
8,950

-110
0
200
-246

Total

93,652,354

93,035,912

616,442

811,872

806,204

5,668

36,658

36,813

-155

Table 10.4 Effect of PAC on electricity generation, fuel consumption and CO2 emissions of the central
production units
10.5 Summary
The four scenarios have been simulated using the techno-economic simulation model and the
scenario outcomes were presented and analyzed to ascertain the balance between environment,
economics and security of supply for each scenario. The results were then compared as a starting
point for the conclusions and recommendations in Chapter 11. Finally, the workings and effects of
electricity storage on the electricity load profile were explained to some degree in order to show some
of the possibilities of the simulation model.

288

References

References
Algemene Rekenkamer (2007), Tariefstelling stadsverwarming [Tariff position district heating]
ANP Pers Support (2005), Amercentrale en Clauscentrale blijven langer in bedrijf [Amer and Claus
power plants to stay operational longer]
ANP Pers Support (2005), Essent overweegt belangrijke upgrade Clauscentrale [Essent considers
major upgrade for Claus power plant]
ANP Pers Support (2005), Essent wil electriciteitscentrale Amer 7 weer in gebruik nemen [Essent
wants to recommission power plant Amer 7]
Brabantsdagblad (2006), Megaclaim tegen Nuon en Essent, 20 January 2006 [Mega claim against
Nuon and Essent]
Budget Net (1999), Electrabel - overname EPON, 22 November 1999 [Electrabel takeover EPON]
Concerted Action on Offshore Wind Energy in Europe (CA) (2001), Offshore Wind Energy, Ready to
Power a Sustainable Europe, Final Report, Delft University of Technology, Duwind 2001.006,
December 2001
CBS (2007), Duurzame energie in Nederland 2006 [Renewable Energy in the Netherlands 2006].
Cogen Projects (2004), Belichten met electriciteit uit eigen WKK of inkopen op de vrije markt, CP
03.136 EK [Lighting with electricity from our own CHP or bought in on the free market?]
Cozijnsen, J., (2006), CO2-handel neemt toe; miljardenmarkt in ontwikkeling, Energiebeurs bulletin
2007/7, August 2006 [CO2 trade grows; billion euro market developing]
CPB (2005), Vier vergezichten op Nederland, Productie, arbeid en sectorstructuur in vier scenarios
[Four Futures of the Netherlands; production, labour and sectoral structure in four scenarios until
2040]
CPB (2006), Welvaart en Leefomgeving, een scenariostudie voor Nederland in 2040 [Prosperity and
Environment: a scenario study for the Netherlands in 2040]
DTE (2007), Meetcode Elektriciteit, Voorwaarden als bedoeld in artikel 31, lid 1, sub b van de
elektriciteitswet 1998 [Electricity Measuring Code; provisions as understood by Section 31.1b of the
Electricity Act 1998]
ECN ( 2001), Buggenum naar Nuon, Energie Verslag Nederland [Buggenum goes to Nuon]
ECN (2004), Potentials and costs for renewable electricity generation, a data overview, ECN-C-0-3006
ECN, (2005), Inzet van biomassa in centrales voor de opwekking van electriciteit, ECN-C05-088
[Deployment of Biomass in Plants for the Generation of Electricity]
ECN (2005). Referentieramingen energie en emissies 20052020, [Reference Projections for Energy
and Emissions 20052020].
ECN (2005), WKK in de referentieramingen 20052020, [CHP in the Reference Projections for 2005
2020].
Electrabel (2003), Jaarverslag 2002 [Annual Report 2002]
Electrabel (2006), Electrabel overweegt de bouw van drie nieuwe energiecentrales in Nederland,
Press News [Electrabel considers building three new power plants in the Netherlands]

289

References

EnergieNed (2005), Tariefadvies voor de levering van warmte aan Kleinverbruikers 2006 [Tariff advice
for the delivery of heat to small consumers 2006], 2005-58
Energy Information Administration EIA (2007), Annual Energy Outlook 2007, with projections to 2030,
DOE/EIA-0383 (2007)
E.ON Benelux (2006), E.ON kiest Maasvlakte voor bouw grote electriciteitscentrale, Press News
08.03.2006 [E.ON chooses Maasvlakte as building site for large power plant]
E.ON Netz (2005), Wind Report 2005
EPZ (2001), Jaarverslag 2000 [Annual Report 2000]
Essent (2002), Jaarverslag 2001 [Annual Report 2001]
Essent (2004), Jaarverslag 2003 [Annual Report 2003]
Essent (2004), Milieu overheidsjaarverslag 2003 Vestigingsplaats Geertruidenberg Amer en
Dongecentrale [Government environmental report for 2003: Geertruidenberg Amer and Donge power
plant sites]
Financiele Telegraaf (1999), Epon in handen van Belgisch Electrabel, 19 November 1999 [Epon in
hands of Belgian Electrabel]
Gasunie Trade and Supply B.V. (2006), Jaarverslag 2005 [Annual Report 2005]
International Energy Agency IEA (2001), Wind Energy Annual Report 2000, May 2001
Lako, P., Seebregt, A.J. (1998), Characterisation of Power Generation Options for the 21st Century:
report on behalf of Macro task El, C-98--085
Landelijk Afvalbeheerplan 20022012 (2004), Deel 1 Beleidskader. Gewijzigde versie van april 2004
[National Waste Management Plan 2002-2012. Part 1, Policy Framework, Revised, April 2004].
Landelijk Afvalbeheerplan 20022012 (2004), Deel 2 Sectorplannen [National Waste Management
Plan 2002-2012. Part 2, Sector Plans].
Landelijk Afvalbeheerplan 20022012 (2004), Deel 3 Capaciteitsplannen Gewijzigde versie van april
2004 [National Waste Management Plan 2002-2012. Part 3, Capacity Plans. Revised, April 2004].
Lapuerta, C., Moselle, B. (2001), Recommendations for the Dutch Electricity Market, The Brattle
Group, Ltd.
Ministry of Economic Affairs (2005), MEP subsidiebedrag
E/EP/5725616, 22 December 2005 [MEP Subsidy Incineration Plants]

afvalverbrandingsinstallaties,

Ministry of Economic Affairs (2007), Een onderzoek naar de ontwikkelingen rond de regeling
Milieukwaliteit Elektriciteitsproductie (MEP), Rapport AD/6100991 [A Study of the Developments
around the MEP Scheme]
Ministry of Economic Affairs (2007), Overdracht beheer 110/150 kV-netten, [Transfer 110/150 kV
grids], EZ ET/EM/7142756, 30 November 2007
Ministry of Economic Affairs (2008), Energierapport 2008 Integrale energievisie op grond van de
Electriciteitswet 1998 en de Gaswet, Juni 2008 [Energy report Integral energy vision base don
Electricity Act and Gas Act]
Nieuwsbank (2003), Nuon koopt elektriciteitscentrales van Reliant, 19 February 2003 [Nuon buys
power plants from Reliant]

290

References

Novem (2004),Statusdocument Bio-energie 2004 [Bioenergy Status Document 2004].


NRG (1999), Belgen kopen stroombedrijf Epon voor 6 mld, 18 november 1999 [Belgians buy Epon
electricity company for 6 billion]
Nuon (2004), Jaarverslag 2003 [Annual Report 2003]
Nuon (2005), Jaarverslag 2004 [Annual Report 2004]
N.V. Samenwerkende elektriciteits-productiebedrijven, SEP (1996). Elektriciteitsplan 19972006
[Electricity Plan 1997-2006].
N.V. Samenwerkende elektriciteits-productiebedrijven, SEP (1996). Toelichting Elektriciteitsplan
19972006 [Explanatory Notes to the Electricity Plan 19972006].
Profundo (1999), PreussenElektra wil ook een distributiebedrijf, November 1999 [PreussenElektra
wants a distribution company as well]
Projectgroep Transitie Biomassa (2003), Biomassa: de groene motor in transitie. Stand van zaken na
de tweede etappe [Biomass: Green Motor in Transition: state of affairs after the second stage].
SenterNovem (2005), Actieplan Biomassa. Statusdocument Bio-energie 2004 [Biomass Action Plan.
Bioenergy Status Document 2004].
SenterNovem (xxxx), Implementatie bio-energie in 2010. Perspectief per Biomassa Technologie
Combinatie. Openbare samenvatting [Implementation of Bioenergy in 2010. Perspective per Biomass
Technology Combination. Public Summary].
TenneT (2000), Capaciteitsplan 2001-2007 [Capacity Plan 2001-2007]
TenneT (2002), Capaciteitsplan 2003-2009 [Capacity Plan 2003-2009].
TenneT (2004), Toelichting import en export van elektriciteit, CA 04-0056 [Notes on the Import and
Export of Electricity]
TenneT (2006), Transportbalans 2006 [Transport on Balance 2006]
TenneT (2006), Rapport Monitoring Leveringszekerheid 2005 2013, MR 06-231 [Report on
Monitoring Security of Supply 2005-2013]
UCI (2006), The Economics of Nuclear Power, Briefing Paper, 8 April 2006
Van Eck, T. (2007), A New Balance for the Energy Sector: no longer a puppet in the hands of
technology, public interests and market, ISBN 978-90 788 8901-4
Van Eck, T., Rdel, J.G., Verkooijen, A.H.M. (2004), Trading between countries improving the balance
between Environment, Economy and Security of Supply or power play? The EU situation, paper
presented at 19th World Energy Congress, Sydney, Australia, September 5-9, 2004
Werkgroep Afvalregistratie (2004) Afval verwerking in Nederland: gegevens 2003 Vereniging
Afvalbedrijven [Waste Processing in the Netherlands: figures for 2003 from the Dutch Waste
Management Association].
Vereniging Afvalbedrijven (2004), Energie uit Afval, Statusrapportage 2004 [Energy from Waste,
Status Report 2004].
Verkaik, J.W. (1999), Evaluation of Two Gustiness Models for Exposure Correction Calculations,
Journal of Applied Meteorology Volume 39, No. 9, pp. 1613-1626, 2000.

291

References

Internet Addresses
1.
2.
3.

292

http://www.enerq.nl/informatie/Tarieven/
http://www.tennet.org/bedrijfsvoering/ExporteerData.aspx
http://www.dte.nl/nederlands/elektriciteit/regelgeving/secundaire_regelgeving/codes/

Conclusions and Recommendations

294

Chapter 11 Overall Conclusions

Chapter 11

Overall Conclusions
11.1

Summary

Much of the current electricity and heat supply system in the Netherlands is based on the generation
technology of the 1980s and 1990s. There has been no new investment in large-scale generation capacity
in recent years. However, investment has increased in capacity that is generated on a more limited scale,
with the emphasis on sustainable wind energy. At the moment, various plans are in the pipeline for new
investment in both large- and small-scale generation.
While these plans were materializing, the Dutch government set ambitious goals for energy and climate
policy in the coalition agreement of 2007. The ultimate aim was to provide the Netherlands with one of the
most sustainable and efficient energy supply systems in Europe by 2020. Basically, the goals are to
realize 2% energy savings every year, to increase the share of renewable energy in the overall supply
to 20%, and to lower greenhouse gases (CO2) by 30% compared with levels in 1990. The deadline for
these objectives is 2020. Substantial investments will be needed to realize such ambitious goals for heat
and electricity, but more importantly, a course will need to be charted that can bring them within reach. The
generation technologies which are chosen now will have implications for at least three decades because
large investments are written off over many years and need to be earned back. Incentive schemes will
continue to be necessary as not all types of capacity particularly renewable capacity will be able to
compete with cheaper options on a deregulated electricity market. It is important to ascertain the costs of a
course in which the principle of choice is pitted against economics and security of supply. It is for this
purpose that the following question was asked in the introduction.
Which scenarios are viable in the Dutch energy system with a view to the trends in supply and
demand, protectionism versus globalization, and the importance of the environment in a fully
liberalized electricity market? How do these trends affect the three objectives of energy
saving/lower CO2 emissions, more economic efficiency, and the continuation of a high level of
security of supply? And what support measures are needed from the government?
To answer these problems a simulation model has been developed which is able to calculate the entire
electricity and heat supply in the Netherlands for a whole year on an hour basis. The model is a technicaleconomic production model which addresses the technological aspects of the energy demand
(electricity and heat) and calculates long-term scenarios for the supply side. In these pre-defined
scenarios the balance between environment, economics and security of supply was assessed. The annual
CO2 emissions and the percentage of electricity generated from renewable (sustainable) sources have
been calculated with a view to the environmental situation. The annual costs of the system and the costs
and investments needed for incentive schemes have been worked out with a view to the economic
situation. Finally, the reserve capacity requirements and how they are influenced by renewable energy
have been calculated for security of supply.
The model makes use of sub modules and a unit commitment/dispatch module to work out the contribution
of specific generation groups in the electricity and heat supply, the fuel consumption and the accompanying
CO2 emissions. The sub modules relate primarily to smaller-scale generation such as CHP, waste
incinerators, and sustainable sources of energy because it is assumed that smaller-scale generation would
contribute to the electricity and heat supply system in predetermined deployment patterns (except
renewables) on the basis of forward pricing for electricity and gas. Larger-scale generation is calculated
within the applicable constraints in a unit commitment / dispatch module on the basis of minimum costs.
The results of these calculations have been aggregated with the results from the sub modules to produce
the outcomes described earlier.

295

The reference year for the generation technology is the electricity and heat situation in 2003. Smaller-scale
generation is defined as blocks with efficiency, heat /power ratio and emissions that were applicable in
2003. In larger-scale generation each unit was modelled separately with heat rate curves, minimum
up/downtime, start-up time etc. In the future, smaller-scale generation will be modelled with improved
generation output etc. For larger-scale generation some 50 new separate units of different types have been
modeled on the basis of anticipated technological developments. These can be deployed according to the
scenario and include carbon capture and storage, and the storage of electricity.
Like every other model, this model has its limitations. The results are based on a transparent and
controllable techno-economic world with an electricity market that functions perfectly and sustains no
impact from the physical chain of fuel supply, production, electricity market, grids and consumers
(chain dependence). The model has first and foremost been used to calculate the costs and not the
returns of the energy supply system. It would be more or less impossible to work out the returns
because of the (many) submarkets (bilateral, APX, imbalance market, breaking capacity, import), each
with their own supply and demand price mechanism. Conclusions about the returns can be drawn only
on the basis of forward commodity pricing but as the relevance of such conclusions is questionable,
returns have not been addressed.
The model calculates and optimizes on the basis of the entire supply system, whereas, in practice,
producers optimize on the basis of their own portfolio and (may) therefore lose out on the synergy
effects.
The influence of wind energy on the supply system will increase as more wind energy becomes
available. Obviously, it is impossible to predict the volume of wind, but a reference year still had to be
selected. The choice is 2001. However, given the influence of wind on the daily operations of the
system, every other year with more or less wind than 2001 will influence the performance of the
system.
The model has been used to calculate long-term viable scenarios for the Dutch energy system. An
inventory have been compiled of the main drivers. This has been used to build a matrix of extremes
which served as a basis for four possible scenarios. In two scenarios, A and C, the emphasis is on the
creation of a sustainable, low-emission energy supply system in a protectionist world (Scenario A) and
a globalizing world (Scenario C). In scenarios B and D, the emphasis rested on the creation of an
energy supply system with the lowest costs in a protectionist world (Scenario B) and a globalizing
world (Scenario D).
Results
Four scenarios have been worked out for the Dutch electricity and heat supply system, starting with the
reference year (2003) and moving toward possible future situations depending on the scenario. Each
scenario has then been assessed for 2012 and 2025 in relation to the environmental, economic and
security-of-supply goals in Chapter 10. The following conclusions can be drawn:
Environment
It has been assumed in all the scenarios that the demand for electricity continues to grow. Because of
energy-saving measures scenarios A and C achieve a growth rate of 1 - 1.5%. In scenarios B and D,
where there have been no energy-saving measures, growth of 2 - 3% has been achieved. Scenarios A and
C have been primarily pro-environment scenarios, but the method of generation has been totally different in
each case. Scenario A makes maximum use of renewables in the form of on- and off-shore wind capacity,
while Scenario C makes much more use of carbon capture and storage and less use of renewables. Both
scenarios, however, use biomass as fuel via the further expansion of waste incineration, the indirect cofiring of clean wood and the gasification of wood waste, agro residues and chicken manure. In addition, the
use of (residual) heat for process heat, district heating and greenhouse heat has been far more intense in
both scenarios, thereby avoiding the use of primary fuels and reducing CO2. Industrial CHP capacity for
process heat has been also substantially expanded but this can, of course, also be realized in large-scale
generation provided the generation site (near the customers) is properly defined. Electricity import in
scenarios A and C has been limited to sustainable electricity only. This makes an absolute reduction of
more then 30% in CO2 emissions realizable in both scenario A and C compared with the reference
situation in 2003. Almost 30% of the energy comes from sustainable sources in scenario A and almost
20% in scenario C. In scenarios B and D the development of renewables has been halted and CHP is

296

Chapter 11 Overall Conclusions

consolidated. In scenario B, where there is a relatively high proportion of nuclear energy, CO2 emissions
rose by 10% in absolute terms. In scenario D, where there is a relatively high proportion of coal capacity
(without CCS), CO2 emissions rises by a startling 63% in absolute terms.
Economics
Environmentalism comes at a price. This hypothesis has clearly been confirmed by the specific overall
costs for scenarios A and C. Due to higher total investment costs, shorter depreciation periods and higher
O&M costs, the annual specific overall costs for these scenarios are higher than for scenarios B and D.
Scenario C fared better than scenario A: though the annual costs for scenario C are higher, more electricity
is produced which loweres the overall costs. It should be mentioned here that the estimated investment in
maintenance costs for CCS was conservative. In other words, if this technology is to develop quickly and
on a large scale (learning curve) it shall lead to lower specific overall costs for scenario C in particular and
push up its score even further compared with scenario A. Scenario B, with a relatively large percentage of
nuclear power, scores best for specific overall costs, though scenario D is a close second. The annual total
costs for scenario B are, however, much lower than for D. The pro-environment scenarios A and C present
a specific rise in cost price of between 7 and 13 /MWh compared with scenarios B and D. Extensive and
consistent incentive schemes will be required for both scenarios. Scenario A with an annual incentive
budget of 1.6 billion is by far the most expensive, followed by scenario C.
Security of Supply
Assuming that it will always be possible to supply electricity, a large volume of reserve capacity will be
needed, especially in scenarios A and C, where the deployment of renewables is relatively high. Scenario
A has by far the worst score for security of supply. Given an installed capacity of 8,000 MWe wind and
1,500 MW solar energy and assuming that 20% of this may count as available capacity, approximately
7,500 MW will have to be held in reserve to compensate for shortfalls. For scenario C this works out at
approximately 4,250 MWe. The swing in the generation capacity of wind in particular presents an additional
problem because it will have to be offset by the rest of the generation system. The lower efficiency of units
in partial operation will lead to extra fuel consumption. In scenario A, where there is an annual increase of
1% in electricity use, the supply of wind capacity will need to be lowered to avoid the non-operation of other
larger-scale generation or the industrial and small scale CHP as this could adversely affect security of
supply. This problem will only get worse if electricity is imported. One storage basin (PAC island) with a
loading/unloading capacity of 1,500 MW would offer some consolation but is nowhere near enough to
absorb the potential swing of scenario A. The effect of one storage basin is more apparent in scenario C.
Scenarios B and D need only a very limited amount of reserve capacity to absorb periods of non-delivery
caused by the maintenance of one large unit and a disturbance in another large unit.
Balance between the Environment, Economics and Security of Supply
The results show that the objectives for the environment, economics and security of supply cannot be
optimally achieved in one scenario. Environment scores well in scenarios A and C, but at the expense of
economics and security of supply. The situation is reverse for scenarios B and D. However, bearing in mind
the level of realism in each scenario, a credible picture does emerge when all four are compared.
Scenarios A and D are not very realistic as maximum wind capacity is assumed in scenario A and the
cheapest generation methods are assumed in scenario D. Scenario C is more realistic as it aims for an
optimal environmental result but with a balanced mix of production technologies comprising renewables,
residual heat and CCS. Again, this scenario needs a large reserve capacity because of the share of wind
energy, but this problem can be solved by extending the CCGT capacity. That said, this capacity can no
longer be efficiently used to deliver heat because of the unpredictable nature of wind energy and the low
number of operational hours. A storage basin with capacity for 1,500 MW could make up part of the CCGT
shortfall. Another possible solution is to replace part of this gas-fired capacity with small- or larger-scale,
well regulated nuclear power from scenario B, where the deployment of natural gas is limited as much as
possible to high-quality CHP applications instead of only electricity from CCGT.

297

11.2

Summarizing Conclusions

The model shows the effects of four different scenarios on investments at national level, the annual
costs of the energy supply, and the costs of lowering/removing CO2 emissions (price/quality ratio).

The instruments in the model (CO2 tax, MEP-type compensation system, energy tax on fuel input)
enable the effect of government intervention on the total costs and on the CO2 emissions to be
measured.

The effects of technological developments, the shift from central to local capacity, the influence of
demand side management etc. can all be modeled.

It is possible in scenarios A and C to achieve a 30% reduction in CO2 emissions compared with
the reference situation in 2003 but this will require a major effort from the larger and smaller
providers alike who will have to expand their production capacity.

The combined deployment of renewables, residual heat for process heat, district heating and
greenhouse heat, and CCS in scenario C is least expensive in environmental terms and offers
potential for further savings if CCS technology gets off to a good start.

A low-emission energy supply system will push up the costs by approximately 10 /MWh
compared with the lowest cost scenarios B and D. These costs will be borne by society
(individuals, industry, government).

The effects on security of supply do not bode well, if all the efforts are concentrated on wind
energy combined with low growth in electricity use, e.g. 1% in scenario A.

A large percentage of wind energy, when combined with low growth in the use of electricity, will
limit the possibilities of residual heat delivery as the unpredictability of wind energy will reduce the
operational hours of other capacity.

Given the target of a 30% reduction in CO2 emissions, scenario C comes out best for the balance
between the environment, economics and security of supply. Further improvement can be realized
by accelerating the deployment of CCS technology, by introducing a storage basin and possibly by
replacing CCGT reserve capacity with a limited amount of small-scale, well regulated nuclear
capacity.

11.3

Recommendations

Now that the government has formulated clear aims for energy saving, sustainable energy and the
reduction of CO2 emissions in the coalition agreement, choices need to be made which will ensure that
the energy supply system reflects a healthy balance between the environment, economics and
security of supply. It should be borne in mind that choices which are made now for specific types of
generation will have implications for the next 35 years. It is therefore essential to define a clear
framework within which the chosen types of generation can develop and deliver the objectives by
2020 or 2025. This must take place in such a way that the trend toward a sustainable/low-emission
energy supply system can continue after 2025.
A simulation model that is developed for this research worked out the implications for the environment,
economics and security of supply in four pre-defined scenarios. Scenario C is particularly successful in
meeting the objectives and in achieving a reasonable balance. It would be advisable to develop
scenario C further after receiving input from the government, the energy providers and the NGOs. The
model can also provide insight into the type of framework within which certain types of generation can
grow. Further, incentive schemes can be mapped out till 2025 and can therefore be consistently linked
to the desired volume of capacity.

298

Chapter 11 Overall Conclusions

The model is currently geared to the electricity and heat supply system in the Netherlands but it can
be used for any country in the EU (and elsewhere) provided sufficient data on supply and demand are
available. It can also be further developed to include the energy supply systems in neighbouring
countries so that the generation framework can be defined at a more European level and meet the EU
objectives for environment, economics and security of supply.
The model is well suitable for analyzing other scenarios in terms of the effects on the environment,
economics and security of supply. It can determine whether the selected energy supply system is
workable and whether the electricity market can continue to function effectively. A realistic simulation
of the entire generation park is achievable thanks to the hour-based modeling and the separate
modeling of all larger-scale generation. This and the easily adaptable input variables mean that a
highly realistic picture can be drawn of the future.

299

300

Appendices

302

Appendix A CO2 Emissions of Imported Electricity for 2003

Appendix A

CO2 Emissions of Imported Electricity for 2003


Netherlands 2003

Physical electricity exchanges 2003

Germany, Electricity imported


Germany, Electricity exported
Germany, Electricity imported net

GWh
GWh
GWh

15,038
601
14,437

France, Belgium Electricity imported


France, Belgium Electricity exported
France, Belgium Electricity imported net

GWh
GWh
GWh

5,779
3,212
2,567

CO2 emissions imported Germany

Mton

8.7

CO2 emissions imported France*

Mton

0.2

CO2 emissions imported Total

Mton

8.8
Source: UCTE, 2003

Germany 2003
Gross electricity production
Net electricity production
Distribution losses + own consumption

TWh
TWh
TWh

607.4
524.4
83.0

CO2 emissions

Mton

315.6

CO2 emissions net electricity production

Mton/TWh

0.602

CO2 emissions imported Germany

Mton

Coal
Lignite
Oil
Natural Gas
Nuclear
Wind
Hydro
- of which regenerative
Other Fuels
- of which Biomass
- of which Solar PV
- of which Waste
Total

8.7

Electricity
TWh
146.5
158.2
9.9
61.4
165.1
18.7
24.2
19.0
23.5
6.5
0.3
4.3
607.4

Fuel
PJ
1,389
1,539
73
410
1,802

5,213

E Efficiency
%
38.0
37.0
48.9
53.9
33.0

CO2 factor
kg/GJ
94.6
101.2
74.1
56.1
0.0

CO2
Mton
131.4
155.7
5.4
23.0
0.0

315.6

France
Gross electricity production 2004
Net electricity production 2004
Distribution losses + own consumption

TWh
TWh
TWh

CO2 emissions net electricity production

Mton/TWh

CO2 emissions imported France

Mton

Coal
Oil
Natural Gas
Biomass
Waste
Nuclear
Hydro
Geothermal
Solar PV
Solar Thermal
Other Sources

572.2
546.7
25.5
0.06
0.2
Share
%
5.0
1.0
3.2
0.3
0.6
78.3
11.3
0.0
0.0
0.0
0.2

CO2
Mton/TWh
0.897
0.546
0.375
0
0
0
0
0
0
0
0

*) Electricity generation in France is taken for calculation of CO2 emissions

303

304

Sources:

Waste

32,471 TJ

Others

66,524

4,887

Hot Windbox CCGT

Own Consumption

Others

Net Import

178,605

146

7,191

16,995 GWh

30,246 GWh

Excelsheet : Overzicht 1999 - 2004 Elektriciteit produktiebedrijven Rev1.xls [Summary 1999 - 2004 Electricty Production companies]

12,261 GWh

98,938 GWh

63,959 GWh

LEI; Energie in de glastuinbouw van Nederland, Ontwikkelingen in de sector en op de bedrijven t/m 2003 [Energy in Horticultural sector in the Netherlands, Developments in sector and companies till 2003]

1,370 GWh

30,246

236

2,040

1,184

1,626

Non Fossil
Waste Incinerators

26,334

1,890

141,860

4,910

967

20,466

Production
Electricity
Heat

Small CHP (Gas engines)

CCGT DH

Industrial CHP

31,616

2,565 GWh

Other Electricity Producers

Own Consumption

31,224

17,187

7,751

CCGT DH

64,416

10,607

4,185

1,411
10,250

61

3,430

CCGT CHP

CCGT

IGCC

Peak / Blackstart

4,018

6,588

Conventional Gas

Nuclear

25,265

Production
Electricity
Heat

Conventional Coal

Electricity Production Companies

www.mnp.nl/mnc/i-nl-0013; Energieverbruik in land- en tuinbouw 1990 - 2003 [Energy use in Agricultural and Horticultural sector 1990 - 2003]

Annual environmental reports individual generation units 2003

EnergieNed; Energie in Nederland 2005 [Energy in the Netherlands 2005]

CBS, Duurzame energie in Nederland 2004 [ Renewable energy in the Netherlands 2004]

397,721 TJ

53,318 TJ

33,142 TJ

9,451 TJ

348,232 Ton
14,297 TJ

3
9.1 billion m
287,513 TJ

572,063 TJ

7,127 TJ

9,211,691 Ton
230,580 TJ

37,512 Ton
1,549 TJ

3
8.2 billion m
259,072 TJ

Biomass

20,804 GWh

17.3 billion m

41,264 TJ

Import

92.5 billion m

Uranium

Coal

Oil

Natural Gas

CBS Statline; Productiemiddelen Electriciteit [Production options electricity]

1,009 GWh

8,704 GWh

11,091 GWh

68.3 billion m

24.1 billion m

Energy Balance-sheet for Reference Year 2003

Appendix B

1,444 GWh

2,856 GWh

15,199 TJ

3,948 TJ

3
2.5 billion m

57,751 GWh

9,405 GWh

173,520 TJ

9,762 TJ

3
11.3 billion m

35,407 GWh

7,400 TJ

3
15.8 billion m

29.6 billion m

Export

Grid Losses

Green Houses

Large Consumers

gas 6,469,000
elek 7,254,000
heat 249,000

Small Consumers

3,809 GWh

4,336 GWh

45.6 billion m

305

Appendix B Energy Balance-sheet for Reference Year 2003

306

Appendix C Description of the PowrSym3

TM

Computer Model

Appendix C

Description of the PowrSym3TM Computer Model (Source: OSA)


Introduction
PowrSym3 is a multi-area, multi-fuel, chronological production cost simulation model for electric
power systems or combined heat and power systems. It has the accuracy and level of detail
necessary for short term operational studies as well as long term planning studies. The detailed
simulation algorithms are combined with a user-friendly database and reporting system to create a
simulation model with a wide range of features, accurate simulations, and a very flexible user
interface.
PowrSym3 is the result of a joint development effort by Operation Simulation Associates, Inc. (OSA)
of the USA and The Netherlands Utility, Sep. Support has also been provided by the Tennessee
Valley Authority (TVA) of the USA.
Key features of PowrSym3 are:

Preservation of chronology
Combined heat and power optimization
Unit commitment with dynamic optimization
Energy storage simulation
Multi-area simulation
Fuel contract model with optimization
Monte Carlo forced outage model
User-friendly database system with flexible reporting

PowrSym3 Overview
General Model Description
The purpose of this section is to give a brief overview of PowrSym3.
The execution sequence of PowrSym3 is shown in Figure C.1. There are three execution time
horizons in PowrSym3: annual, monthly, and weekly. The annual horizon is used for reliability
calculations, adjusting expansion unit installation dates, and maintenance scheduling. The weekly
horizon is used for production cost optimization. The basic operational time step is one hour.
The control file links input and output files to specified file names. Input data is updated at the
beginning of each execution year and may change on weekly boundaries (monthly changes with
monthly optimization). Additionally many input items are allowed to change hourly.
The annual reliability model computes the annual loss of load probability (LOLP) in hours per year
using the method of moments (or cumulants). Load carrying ability and capacity surplus/deficit and
desired reserves are calculated relative to a specified reliability index.
The expansion module optionally schedules capacity additions to the system to satisfy capacity
deficits. These additions may be made in a pre-determined sequence from an input list or the logic
will choose from base load, intermediate load, or peaking options based on system energy
requirements. This module is not a substitute for optimal system expansion studies, but allows
PowrSym3 to schedule pre-selected expansion stations as needed to meet system reliability
requirements.

307

P ow rS ym 3

STA R T

E x e c u tio n F lo w
R e a d C o n tro l F ile

R e a d In p u t D a ta fo r 1 y e a r

F in is h e d ?

Y es

N o

End

A nnual
R e lia b ility C a lc u la tio n
S e le c t
E x p a n s io n U n its

D e te rm in e A n n u a l
M a in te n a n c e S c h e d u le
P rin t
A nnual
LO LP
R e p o rts

Y ear
C o m p le t e
P rin t
A nnual
R e p o rts

E x e c u te
B y W eek
?

N ext W eek
S e le c t M o n te C a rlo
O u ta g e D ra w s

N ext
D raw

E x e c u te
B y D raw s
?

C o m p le te
O u tp u t
R e p o rts
W e e k ly S im u la tio n

Figure C.1: PowrSym3 Execution Flow


The annual maintenance schedule can be determined in one of two ways. The schedule may be
determined external to PowrSym3 and input through a file or PowrSym3 may determine the
schedule. In the internal schedule option, the scheduling objective function is to levelize weekly
LOLP.
There is also a single-pass Monte Carlo outage draw method optionally implemented by the
maintenance schedule module. This method, known as the semi-guided method, draws the expected
amount of outage hours during the year for each station randomly selecting the hours, but using some
guidance to distribute the total amount of outages evenly (within bounds) over each week of the year.
This method, while not statistically converged, is very good for long range planning studies.
The first step in weekly execution is to determine the weekly random outage draws, unless
predetermined by input data, the semi-guided method or deratings. The Monte Carlo logic, either by
random draws or the smart draw method, selects the hourly station outage states for a specified
number of iterations. The random method selects states driven by a random number generator. The
smart draw method generates a large number of iterations and then uses a very fast production cost
approximation to select a sample set of iterations to represent the entire distribution.
308

Appendix C Description of the PowrSym3

TM

Computer Model

Each iteration is then simulated and the results added to the output arrays. Both the expected value
across all iterations and the value of each individual iteration are available for output.
The weekly simulation, shown in Figure C.2, schedules the hourly operation of each resource over the
weekly execution horizon. The hydro stations are scheduled first using a load leveling algorithm.
Each hydro station is subject to a weekly energy constraint, hourly minimum and maximum generation
levels, and ramp rates.

P o w rS y m 3

ST A R T

W e e k ly S im u la tio n
S c h e d u le H y d ro fo r
w e e k b y h o u rs

I n i tia l T h e rm a l
S im u la ti o n

S c h e d u le E n e rg y
S to ra g e

T h e rm a l
S i m u la tio n

T h e rm a l
S im u la ti o n

G e n e ra te
D P C ases

For
E a c h D P C a se
?

N ex t D P C a se

C o m p le te
T h e rm a l
T h e rm a l
S i m u la tio n S i m u la tio n

D P C o m m it
O p t im iz a tio n

If d e sir e d

F IN A L
T h e rm a l
S im u la ti o n

O u tp u t
R e s u lt s

End

p s-

Figure C.2:- PowrSym3 Weekly Simulation


The thermal simulation schedules hourly thermal plant operation to meet electrical and heat loads.
The initial thermal simulation is run without energy storage operation and creates a multi-step hourly
marginal cost array. The marginal cost array is used as the objective function for scheduling hourly
energy storage operation in the next step. The thermal simulation is then executed again with the
given hourly energy storage schedule.
As an option, a dynamic programming module can be used to further improve the unit commitment
schedule. An expert system, driven by user input rules, can then used to generate a series of
additional thermal simulations with variation in the thermal unit hourly commitment. A dynamic
programming algorithm is used to select the hourly commitment from the available simulations with an
objective function of minimizing total system cost over the weekly horizon. A final thermal simulation is
then made using the unit commitment determined by the dynamic programming module.

309

PowrSym3 has two thermal station simulation methods available, the Hourly Sequential Method and
the Concurrent Simulation Method. The two methods are interchangeable except that the fuel contract
logic is functional only with the Concurrent Method. The Hourly Sequential Method has evolved over
several years while the Concurrent Method is new and still rapidly evolving. It is expected that the
computational advantages of the Concurrent Method will result in it becoming the method of choice.
Input Data Organization
The PowrSym3 Data Flow is shown in Figure C.3. Database file can be prepared directly using a
text editor or by using a database as described below.

Power
System
Data

Database

On Line
Help

Hourly
Load
Data

Weekly
Maintenance
Data

PowrSym3

Example:
MS ACCESS

Reports

Hour

Week

Month

Calendar
Year

Fiscal
Year

Figure C.3: PowrSym3 Data Flow Diagram


The most user friendly input method is the use of a database product such as Microsoft ACCESS.
The PowrSym3 data file is in a format, which is easily imported and exported by any database or
spreadsheet product. The advantage of using the same file format for import and export to the
database is that changes made outside of the database with a text editor can easily be brought back
into the database.
Database templates for use with ACCESS are provided with PowrSym3. These templates provide
context sensitive help, on-line indexed help, and numerous forms, queries, and macros to assist in
data preparation, verification, and reporting.
Output Data Organization
The output from PowrSym3 can be written into a number of output files that the user can print and
review. Alternatively the user may choose to have the output written into a spreadsheet or database
format where it can be viewed and manipulated by commercial models for presentation.

310

Appendix C Description of the PowrSym3

TM

Computer Model

The output routines produce results for the following quantities by generating station:

Capacity factor
Fuel burn (Equivalent energy and physical units such as tons or barrels)
Costs (fuel; operation and maintenance, O&M; start up; emissions; total)
Emissions (five user defined categories)
Electric energy production
District heat energy production
Industrial heat energy production
Number of unit starts
Operating hours

Results are summarized by fuel type and optionally may be summarized by station, plant, and area or
power pool. Output reports, by default, summarize the above quantities by month and fiscal year with
options for calendar year, weekly, and hourly reports. Optionally PowrSym3 can also produce hourly
marginal cost reports.
An annual report lists for each year the following items by week:

LOLP in hours
Peak and average loads
Load factor
Installed capacity before and after maintenance
Percent reserve before and after maintenance
Capacity on maintenance

The fuel contract report lists the quantity burned and purchases cost of each fuel. Any required fuel
purchased that could not be utilized because of outages or other system constraints is listed also. An
optional report shows the sources and amounts of fuel by station and the distribution to stations by
fuel contracts.
The TRANSAREA report lists the load, generation, imports, and exports of each area. The link report
lists the energy and maximum transfer over each link.
Annual Reliability Model
PowrSym3 has an integrated loss of load probability algorithm, which computes system reliability
using probabilistic techniques. The actual reliability and the surplus or deficit of system capacity
relative to a specified reliability index is computed. The model will optionally adjust the installation
dates of new capacity additions in a specified sequence to meet system reliability requirements. Loss
of load probability and expected unserved energy are computed hourly and summed by week, month
and year Capacity requirements may be computed relative to a weekly index, and annual index, or
both.
The Annual reliability Model does not recognize the detailed hourly capacity and unit availability
variations, which are possible in the production cost simulation and in Weekly Reliability Model.
Instead each resource has a separately specified dependable capacity which is used in the Annual
Reliability Model and in the Maintenance Schedule Model. The dependable capacity of each resource
may vary seasonally. The requirement to have all data in memory for the entire year prohibits the use
of hourly variation patterns.
Maintenance Schedule Model
Pre-scheduled Maintenance
The pre-scheduled maintenance is read from a file which contains start and end date of each outage.
The dates are specified by week and hour of week so that each unit outage can be scheduled down to

311

the hourly level. A generating unit may multiple outages during the year. The maintenance algorithm
executes in fiscal year increments so an outage may not overlap the fiscal year boundary. An outage
that overlaps the fiscal year boundary, must end at the end of the fiscal year and a new outage begin
at the beginning of the following fiscal year A user-friendly interface is provided for building the
maintenance file.
The maintenance schedule file may be prepared manually from an existing or planned schedule or
may be prepared by automated maintenance scheduling model or saved from a previous PowrSym3
run.
Calculated Maintenance Schedule
PowrSym3 can estimate a future maintenance schedule based on levelizing the weekly loss of load
probability. planned maintenance outage rate of each generating unit is translated into a single
contiguous outage of the appropriate number of hours. The maintenance events are then scheduled,
large units first, with the objective function of levelizing weekly loss of load probability and minimizing
the annual loss of load probability.
Maintenance rules are invoked by the use of maintenance groups. Each generating unit may belong
to several maintenance groups. Examples of rules that may be invoked for maintenance groups are:

No more than N units of this group out simultaneously.


No more than X mw of this group out simultaneously.
Minimum of N hours between the end of a maintenance and the beginning of a new
maintenance in this group.

The reliability index (LOLP) is computed as the objective function for maintenance scheduling and is
also included in the output reports. The index is computed as the expected hours per week of
capacity deficiency. The index is computed for each hour of the week independently and summed for
the week. The computational method is based on a simplified version of the cumulant method and
does not go to the level of detail of Weekly Reliability Model. However, the results track with those of
more detailed models, and the method serves quite well as an objective function for maintenance
scheduling.
The Maintenance Schedule Model does not recognize the detailed hourly capacity and unit availability
variations that are possible in the production cost simulation and in the Weekly Reliability Model.
Instead each resource has a separately specified dependable capacity which is used in the Annual
Reliability Model and in the Maintenance Schedule Model. The dependable capacity of each resource
may vary seasonally. The requirement to have all data in memory for the entire year prohibits the use
of hourly variation patterns.
Combined Maintenance
It is possible to combine an external maintenance schedule and calculated maintenance option. For
each unit, any scheduled maintenance outages will be deducted from the calculated maintenance
outage rate and the remainder, if any, will be scheduled.
Weekly Reliability Model
PowrSym3 has a very accurate reliability algorithm which may optionally be executed weekly during
the simulation. This algorithm uses direct convolution and double precision arithmetic for accurate
calculations.
This algorithm takes into account hourly variations in unit capacities and availability for a more detailed
reliability calculation than is available with the Annual Reliability Model. Output reports include hourly
unserved energy, and a weekly summation.

312

Appendix C Description of the PowrSym3

TM

Computer Model

Forced Outage Model


The PowrSym3 chronological production costing model has six modes simulation for forced outages.
All of these modes can use the same data set. The six modes of simulation are:

Fixed outage schedule


Derating method
Gradient derate method
Random Monte Carlo
Selected Sample Monte Carlo
Semi-guided Monte Carlo

The representation of forced outages and forced deratings in chronological production costing models
has been the subject of much research in recent years. The six methods available in PowrSym3
represent most of the methods available in the literature. Each of the methods has advantages and
disadvantages for various types of studies. Unfortunately the perfect solution awaits new methods or
faster computers.
Fixed Outage Schedule
Forced outages and forced deratings may be input to PowrSym3 as a given input, and the model will
assume all remaining resources to be available with certainty. This method is suitable for very shortterm studies of specific scenarios. This method is fast because no iterations are required. The fixed
outage schedule is often used for debugging and model comparisons in order to remove the
fluctuations caused by varying outage scenarios.
Derating Method
The derating method reduces the capacity of each generating unit by the forced outage rate. For a
large system, this is approximately the same as the statistical method used in older versions of
POWRSYM. This method is fast because no iterations are required, but the operation of peaking
resources such as pumped hydro and combustion turbines will not be accurate. This method tends to
understate the use of pumped hydro and units while overstating the use of intermediate units. The
amount of capacity on outage is always at the average condition. No extremes are observed.
Derating methods are not recommended for small systems.
Gradient Derate Method
The gradient derate method is a TVA specified method of derating for forced outages in the short term
operational studies. It approaches the standard derating method after several hours are simulated.
Random Monte Carlo Method
The random Monte Carlo method is a brute force procedure to calculate expected system operation of
the system by averaging a number of outage scenarios created by a random number generator. If
enough iterations are used, this is the most accurate method of representing forced outage and forced
derating events but is also the most time consuming. The extreme outage events, both good and bad,
are represented and the operation of pumped hydro and combustion turbine peaking units is realistic.
The outage draw duration is an input to PowrSym3 and may be 168 hours or less. A new set of
random draws will be made at the beginning of each duration period for the state of the units during
the period. A unit is in or out for the duration of the period. With a duration of 168 hours, draws will be
made once per week and with a duration of 12 hours, draws will be made 14 times per week. In a
future version of PowrSym3, the draw duration will also be a random variable for each generating
unit converting the model to frequency and duration mode Monte Carlo.

313

Reducing the draw duration introduces more random draws and speeds convergence with fewer
iterations. However, reducing the duration to much less than the average system outage duration will
increase the number of unit starts and stops above that which will actually occur on the system. A
draw duration of 24 hours is a realistic compromise between accuracy and speed for most systems.
For a large power system it has been found that 50 to 100 random draws are required for
convergence. This means that only one iteration is needed for long range studies in which only the
annual results are of interest, even with a 168 hour outage duration. With a daily duration, seven to
fifteen iterations are required for weekly convergence and two to four iterations for monthly
convergence. With a 168 hour duration, 50 or more iterations are required for weekly convergence.
The Random Monte Carlo is the method of choice short term operational studies. It is often
informative to review the range of outcomes created by the individual draws as well as the average of
the outcomes. These outcomes are often plotted probability curves. It is important to note that the
outcome probability curves are normally shaped for large systems of near equal generating units but
are skewed upwards for most real power systems by the influence of large generating units outages
and can be multi-modal for smaller systems.
Selected Sample Monte Carlo
The selected sample Monte Carlo draw method uses a fast objective function to approximate the
production cost results for each iteration. A large number of iterations, 100 up to 1000, are used to
define the sample space, and the objective function results are chronological constraints relative to
previous hours results and the future hour look-ahead.
The results are sorted to create an objective function probability curve. This curve will not be normally
distributed but will be skewed in the upward direction because of the increasing system lambda
associated with higher outage draws. A smaller number of samples, usually five to eleven and always
an odd number, are then chosen to represent the entire sample space in detailed production cost
calculations. The samples are chosen by dividing the curve into equal areas. This method insures that
a small sample is not overly weighted by a bad draw and that the draws are uniformly selected to
represent the entire sample space. A draw sample of one would use the sample nearest the mean of
the distribution. The reason for using an odd number of samples is to force inclusion of the mean
sample.
The strength of this method is that weekly production cost totals are usually very near the converged
values with greatly reduced computation time. Generation by station type and fuel category is also well
behaved. However, individual station results are not well converged and base load plants may not be
operating at their expected capacity factors on an individual basis, although the average for the entire
group of base load plants is likely to be near the expected values.
A special consideration for the selected sample method is repeatability of the draw sample. When
comparing model results, it is important that each run have the same random draws. Since the model
is selecting the draws to be used in this method, it is necessary to save the draw sample file from the
first run and force the following runs to use the draw sample file instead of creating new draws.
Semi-Guided Monte Carlo
The Semi-Guided Monte Carlo method is described in detail in paper titled Using a Semi-Guided
Monte Carlo Method for Faster Simulation of Forced Outages of Generating Units presented by
Scully, et. al., at the Seventeenth PICA Conference. This paper was reviewed by T. Jackson and C.
Stansberry of TVA. Variations of this method are used in industry models.
This method produces statistically balanced forced outage schedules over an extended time period,
usually one year. Results for shorter time periods within the year are not converged, but the annual
results for both system totals and individual station results are very well behaved. The method is fast
as only one iteration is required. This is the method of choice for medium and long range planning

314

Appendix C Description of the PowrSym3

TM

Computer Model

studies where only annual results are of interest. The Semi-Guided Method is not recommended for
short term operational studies.
Hydro Model
Hydro stations are scheduled to operate in such a manner as to }levelize the load shape to be served
by other stations. Hydro stations are scheduled sequentially over weekly time horizon, subject to
hourly constraints for minimum and maximum generation, and weekly constraints on ramp rates, and
energy. Typically, hydro levelizes the system load; however, a hydro station can be scheduled to
meet the load of a transmission or control area (e.g., when there are transmission constraints
preventing the station from benefiting the system as a whole).
The station is first scheduled to operate at its hourly minimum for all hours, and the load for each hour
is reduced by the amount of this generation. If this schedule is less than the available energy for the
week from the station, the generation is increased by an increment for the hours with the highest
adjusted loads. The loads for these hours are accordingly adjusted downward. Hourly constraints are
enforced during the dispatch process. This process is continued until the total weekly generation for
this station matches the available energy from the station. Interpolation is used on the last increment.
It is possible to specify the daily maximum and minimum energy to be used by a hydro station. If the
sum of the daily maximums is less than weekly energy allocation, a warning is issued, and all
available energy is not used. If the sum of the daily minimums is greater than the weekly energy, a
warning is issued, and hydro scheduling stops when all energy has been used.
Sequential Dispatch and Commit
The Hourly Sequential Thermal Simulation Method is shown in Figure C.4. This method has evolved
over several years and several versions of POWRSYM and PowrSym3 and is a robust chronological
simulation.
P o w r S y m 3

S T A R T

T h e r m a l S im u la tio n
H o u r ly

S e q u e n tia l M

e th o d

?
M

P re d ic t H o u rly
C a p a c ity R e q u ire d

F in is h e d

E x e c u te
B y H o u r
?

o r e

E n d

S e t C O M M IT a n d
D is p a tc h to m a x im u m

D e c re m e n ta l
C o s ts

m in .

S ta tio n

L o a d

G e n e ra tio n

C o m p u te S ta tio n
D e c re m e n ta l C o s ts

m a x .

C o m p u te
D e c o m m it C o s ts

D is p a tc h H e a t a n d
a d ju s t c o s ts fo r h e a t

S e le c t A re a
w ith h ig h e s t c o s t

R e d u c e G e n e ra tio n
in S e le c te d A re a

Y e s

G e n e r a tio n & L o a d
in b a la n c e
?

N o

th e r m - h r .p p

Figure C.4: Thermal Simulation Hourly Sequential Method

315

The first step in the simulation is a look-ahead algorithm that predicts the capacity requirements of
future hours to aid in commit decisions. The simulation then solves the hours in sequence forcing
chronological constraints relative to previous hours results and the future hour look-ahead.
The hourly simulation begins with all the thermal stations committed and dispatched to the maximum
allowed by constraints related to the previous hours operation minimum down time and ramp rate).
The decremental dispatch cost and the decommit cost of each station are then computed. The
decommit cost includes operating costs and start-up costs amortized over the predicted run time. The
heat loads are then dispatched and the decremental costs are adjusted for heat production effects.
The next step is to determine the transarea with the highest incremental cost that can have its
generation further reduced. A network flow transport algorithm is used to schedule flows between the
areas and compute the decremental cost and potential surplus of each area. The area with the highest
decremental cost and the station within that area with the highest decremental cost are then selected.
If the selected stations decommit cost is higher than its decremental dispatch cost, it will be
decommitted, otherwise its dispatch will be reduced by a decrement.
Decommit decisions are subject spinning reserve requirements operating reserve requirements,
minimum down time constraints, minimum up time constraints, station minimum constraints, and heat
load constraints. Dispatch decisions are subject to ramp rate constraints, unit minimum constraints,
station minimum constraints, and heat load requirements.
After completion of the decrement, costs are re-computed and the process repeated until the
generation and load are in balance. The repetitive process will marginal costs for all areas and all
stations together at system lambda. This method is sometimes called the Equal Incremental Cost
Method.
Concurrent Dispatch and Commit
The Concurrent Simulation Method, Figure C.5, derives its name from its simultaneous solution of all
hours in the optimization horizon (currently one week). This simulation method in a chronological
format is new to PowrSym3 and, to our knowledge, is unique in the industry. It offers accuracy and
speed improvements over the sequential method and allows optimal fuel allocation and fuel, energy,
or emission limits to be simulated.
The simultaneous solution of all hours allows more optimal solution of events which span multiple
hours. Examples of events that must be evaluated over a span of hours are unit commitment and
energy limited fuel contracts.
The concurrent method will find more optimal unit commitment schedules for systems where commit
decisions are strongly influenced by start costs or minimum up and down times. The commitment
schedules developed by the concurrent method are comparable to those developed by dynamic
programming optimization methods. The concurrent method uses dynamic programming techniques
inside a network flow algorithm to produce results much faster than the dynamic programming
optimization methods.
The fuel contract logic allows multiple fuels serving multiple stations. A station may have several fuel
contracts plus spot fuel. Several stations may share a fuel contract. Each contract may have station
specific transportation cost and probabilistic derating. Each contract may have hourly, daily, and
weekly minimum and maximum deliveries and storage with minimum and maximum levels, which
change by season. The concurrent method will optimize use of the fuel contracts to minimize total
system weekly production costs.

316

Appendix C Description of the PowrSym3

P ow rS ym 3

TM

Computer Model

S T A R T

T h e r m a l S im u la tio n
C o n c u r r e n t M e th o d

S e t C O M M IT a n d
D is p a tc h to m a x im u m

D e te rm in e
M A R G IN A L F U E L
fo r e a c h s ta tio n

D e c re m e n ta l
C o s ts

m in .

L oad

G e n e ra tio n

C o m p u te S ta tio n
D e c re m e n ta l C o s ts
fo r e a c h h o u r

m ax.

C o m p u te
S ta tio n D e c o m m it C o s ts
fo r e a c h h o u r

D is p a tc h H e a t a n d
a d ju s t c o s ts f o r h e a t
fo r e a c h h o u r

S e le c t R e d u c tio n
w ith h ig h e s t
D e c re m e n ta l C o s t

M a k e R e d u c tio n

N o

G e n e r a tio n & L o a d
in b a la n c e
?

Y es
E n d

Figure C.5: Thermal Simulation - Concurrent Method

Multi-Area Model
PowrSym3 can be used in the multi-area mode to represent transmission areas within a system
and/or to model interconnected systems. Each station is assigned to area and each area may have a
separate load file or a fraction of a total load file. Each area may be connected to any or all other
areas through links. Each link has provision for a maximum transport capability, a transmission service
charge, and a loss factor. The link data can be specified differently for each direction Pre-scheduled
transactions may be specified between areas over a specified link path.
Spinning reserve and operating reserve requirements may be specified for each area, and the reserve
responsibility for pre-scheduled firm transactions may be assigned to either the sending or receiving
end of the transaction.
Economy power flows between the areas are computed during the simulation process using a
transport algorithm (not a load flow). It is assumed that all feasible economy transactions will be made.

317

Spinning and Operating Reserve


PowrSym3 features a detailed model of spinning and operating reserve with a variety of specification
methods and constraints on the reserve contributions of individual generating units. Spinning Reserve
Requirements must be met by un-dispatched capacity of on-line generating units. Operating reserve
includes spinning reserve plus off-line quick-start generating units. Operating and spinning reserve
requirements may be specified for any combination of system, transarea, and control areas Reserve
requirements maybe specified as a constant amount, a percent of load, the largest on-line unit, or
some combination of these amounts.
The contribution of each generating unit to reserves can also be controlled. A non-firm unit does not
contribute to reserves, a firm unit does. A quick-start unit contributes to operating reserves while offline. An upper bound may be placed on a units contribution thus limiting its contribution when partially
dispatched during low load periods. A lower bound may be placed on a units contribution to reserves
effectively preventing the unit from being fully dispatched unless reserve constraints must be violated.
A summary report of spinning reserve violations is produced.
Combined Heat and Power Production (cogeneration)
PowrSym3 schedules the production of both heat and power to meet the hourly load obligations of
both. Some units produce only electricity, some both heat and electricity, and some only heat. The
algorithm schedules operation of all three types so as to minimize total operating costs for meeting
electric and loads in a combined optimization.
The heat load is specified by heat areas. Each heat area may have an hourly varying district heating
load and industrial steam load. Combined heat and power units and boilers may serve either (or both)
the district heat load and industrial steam load in their area.
The fuel input to a combined heat and power station is:
A + B * power + C * power2 +D * Heat + E * Heat2 + F * power * Heat + G * IHeat
where:
A, B, C, D, E, F, and G
power
Heat
Iheat

- input coefficients
- electric production
- low temperature heat production
- industrial steam extraction

There is also a defined operating area for each combined heat and power station which defines the
allowed range of electric operation as a function of heat production and the allowed range of heat
production as a function of electric production.
Emissions Model
PowrSym3 allows for a variety of emission types methods of computing the emissions. Emissions
may be a simple rate based on a percentage of fuel burn at a station or, as in the case of NOX, a
complex equation based on the dispatch level of the generating unit. Optionally, each emission
category may be assigned a penalty cost that will be included in the dispatch and commit optimization.
Emissions are reported by station and by emission category.

318

Appendix C Description of the PowrSym3

TM

Computer Model

Fuel Contract Model


The Energy Limited Fuels (ELF) Module, Figure C.6, when used in conjunction with the Concurrent
Method, directly solves energy limited fuels allocation. This is accomplished by keeping the ELF fuel
allocated optimally as units are dispatched down and decommitted. For a given commit, ELF
allocation will be optimally solved.
The ELF allocation is solved using network flow programming. Fuel constraints and prices as well as
station demand are modeled using a system of nodes interconnected by links. Each constraint has an
associated set of nodes (i.e., weekly constraints for a fuel are assigned a node, daily constraints for
that same fuel are assigned a set of 7 nodes, one per day, hourly constraints for the fuel are assigned
a set of 168 nodes.) Additional nodes are added to the network in order to represent stations, one
node per hour per station. The last node assigned to the system represents both the source and the
sink. All fuel flows through the system links from the source node to the sink node.
Links are used to interconnect the above set of nodes. Following the example started above, for the
given fuel, the source is connected to weekly fuel node (lelfwk), which is, in turn, connected to each
daily node (lelfdy), which are then connected to their respective hourly nodes (lelfhr). Each hourly fuel
node is connected with the hourly station node for all stations using this particular fuel (lstahr). Each
hourly station node is then connected to the sink (ldemand) thereby completing the circuit. In addition
to the links directly associated with a particular ELF, there are spot links which connect the source with
each station hour (lspot), and dump links which connect each ELF hour with the source.
Each link, connecting a node U to a node V, has a mirror link which connects node V to node U. Flow
is scheduled on links, from source to sink, and unscheduled on mirror links, sink to source. Associated
with each link is its cost (costl), flow (flowl), and potential flow (amntl). The cost for a link is
determined by violation cost, which is the amount of fuel flowed with respect to the constraints for that
link and fuel cost, which is the cost of flowing a unit of fuel on the link. The flow for a link is associated
with the link only, not the mirror link, and represents the net amount of fuel that has passed from node
U to node V. The potential flow for a link is the amount of fuel which may be flowed before the cost of
flowing that fuel changes. This potential flow is associated with both links and mirror links.

319

Figure C.6: ELF Fuel Network

Electric Energy Storage Model


PowrSym3 has a detailed simulation and optimization of electric energy storage devices such as
pumped hydro, compressed air storage, and batteries. Operation is scheduled by making a first
simulation pass without storage operation to develop an incremental/decremental cost function for
each hour of the week. Storage operation is then optimized using these hourly cost functions and a
shadow cost for the storage. The final simulation pass then includes the storage operating schedule.

320

Appendix C Description of the PowrSym3

TM

Computer Model

Storage options have the following characteristics:

Minimum generation, if committed.


Maximum generation
Minimum storing rate, if storing
Maximum storing rate
Maximum storage
Amount in storage at beginning and end of week
Generating and storing efficiencies
Storage natural inflow rate (negative for loss rate)

Heat Storage Model


An iterative optimization for heat buffers is underdevelopment. A test version is now available but the
model is not yet ready for general application.
Marginal Cost Reporting
PowrSym3 calculates the hourly marginal costs by making a system re-dispatch after load change is
made. The re-dispatch must continue to meet all system constraints such fuel contracts, reserve
requirements, and heat loads, these constraints often have significant impact on the marginal cost
results.
The marginal cost report contains the marginal cost for each hour of the study period and is designed
for input to a spreadsheet for summarization by period and other computations.
The algorithm freezes the unit commitment, hydro, and pumped hydro for the marginal cost
calculations because allowing these items to vary can create very large marginal costs in some hours
and negative marginal costs in other hours. The costs associated with unit commitments or changes in
the hydro and pumped hydro schedules can not be allocated to single hour load changes.

321

322

Appendix D Heat Rate figures of Units build in PowrSym3

Appendix D

Heat Rate figures of Units build in PowrSym3


12,000

11,000

Heat Rate [kJ/kWh]

10,000

PCC
9,000

IGCC 2GT+1ST
IGCC 1GT+1ST
CCGT 2GT+1ST

8,000

CCGT 1GT+1ST
PBMR

7,000

6,000

5,000
0

100

200

300

400

500

600

700

800

900

1000

1100

1200

Power [MWe]

Figure D.1: Heat Rate figures of units build in Powrsym3 for first sight period 2012

12,000

11,000

10,000

PCC

Heat Rate [kJ/kWh]

PCC CCT
IGCC
IGCC CCT

9,000

IGCC
CCGT
CCGT
8,000

CCGT CCT
PBMR
ABWR
EPR

7,000

6,000

5,000
0

100

200

300

400

500

600

700

800

900

1,000

1,100

1,200

1,300

1,400

1,500

Power [MWe]

Figure D.2: Heat Rate figures of units build in Powrsym3 for second sight period 2025

323

324

Appendix E Possible grow amounts District Heating

Appendix E

Possible grow amounts District Heating

300,000

Number of Small Consumers [-]

250,000

200,000

2003
2012

150,000

2025
2050
100,000

50,000

0
Randstad Noord

Randstad Zuid

Utrecht

KAN gebied

Tilburg / Breda

Twente

Others

Figure E.1: Number of small consumers in several district heating areas

40,000

35,000

Numer of Large Consumers [-]

30,000

25,000
2003
20,000

2012
2025
2050

15,000

10,000

5,000

0
Randstad Noord

Randstad Zuid

Utrecht

KAN gebied

Tilburg / Breda

Twente

Others

Figure E.2: Number of large consumers in several district heating areas

325

553,494

600,000

368,769

2003

90,590

84,398

78,950
81,179

73,322

42,773

26,888

15,890

146,220

72,382
97,642

54,894

56,824

107,548

129,022

91,946

2050

12,187
30,447

100,000

2025

200,321

206,212

2012

66,279

200,000

157,653

226,451

300,000

272,712

358,753

400,000

110,025

Number of house building equivalents [we]

500,000

0
Randstad Noord

Randstad Zuid

Utrecht

KAN gebied

Tilburg / Breda

Twente

Others

Figure E.3: Number of house building equivalents [we] in several district heating areas

25,000

Heatproduction [TJ]

20,000

15,000
2003
2012
2025
2050

10,000

5,000

0
Randstad Noord

Randstad Zuid

Utrecht

KAN gebied

Figure E.4: Heat production in several district heating area

326

Tilburg / Breda

Twente

Others

Appendix F Results Scenario Outcomes : General

Appendix F

Results Scenario Outcomes : General


220,000
211,977

Electricity Consumption [GWh]

200,000

180,000
171,026
Scenario A

160,000

Scenario B

153,643

Scenario C
Scenario D

144,364
140,000

138,099
132,230
126,548
121,060

120,000

110,935
100,000
2000

2005

2010

2015

2020

2025

2030

Figure F.1: Electricity consumption including grid losses

40,000

39,641

37,500
36,192
35,592

Installed Capacity [MWe]

35,000

32,500

32,258
Scenario A
Scenario B

30,000

Scenario C
Scenario D

27,861

27,500

26,069
25,518

25,000

25,068

22,500

20,000
2000

20,306
2005

2010

2015

2020

2025

2030

Figure F.2: Installed total capacity

327

40,000

35,000

33,715

30,000

Installed Capacity [MWe]

26,332

Reference 2003 Central


Reference 2003 Decentral

25,000
21,802
20,685
20,000

Scenario A Decentral

16,902

16,393

Scenario B Central

15,193

13,545

15,000

Scenario A Central

19,290

18,893

Scenario B Decentral
13,790

Scenario C Central
Scenario C Decentral

9,875

10,000

Scenario D Central

9,125

Scenario D Decentral

7,176

7,176

6,761

5,926

5,926
5,000

0
2012

2025

Figure F.3: Installed central & decentral capacity

200,000
183,149
180,000

160,000
142,314

Reference 2003 Central

Electricity Supply [GWhe]

140,000

Reference 2003 Decentral

120,000

Scenario A Central

111,918

Scenario A Decentral
99,882

100,000

Scenario A Import

93,036

Scenario B Central

84,398
75,076

80,000

Scenario B Decentral

73,467

63,154

Scenario B Import

62,193

43,545
40,000

Scenario C Central

55,730

60,000

Scenario C Import

39,712
32,446

32,348

30,740

Scenario C Decentral

28,828

28,712

Scenario D Decentral
Scenario D Import

20,000
2,438

2,438

2012

Figure F.4: Electricity supply central & decentral

328

Scenario D Central

2,438

0
2025

4,877

Appendix F Results Scenario Outcomes : General

400,000
363,307
350,000
308,883
289,818

300,000

260,387

Heat Supply [TJ]

250,000
215,768

209,453

215,760

215,649

215,652

Reference 2003

200,000
Scenario A
Scenario B
150,000

Scenario C
Scenario D

100,000

50,000

0
2012

2025

Figure F.5: Total heat supply

1,600,000

1,406,852
1,400,000

1,210,398
1,200,000

Reference 2003 Central

Fuel [TJ]

1,000,000

Reference 2003 Decentral

927,426
842,403

800,000

806,204

Scenario A Decentral

673,448

653,879
600,000

Scenario A Central

727,475

Scenario B Central

555,348

581,623

540,403

515,622

Scenario B Decentral
Scenario C Central

463,350
400,000

398,561

393,515

394,031

379,269

379,912

Scenario C Decentral
Scenario D Central
Scenario D Decentral

200,000

0
2012

2025

Figure F.6: Fuel supply central & decentral

329

330

Appendix F Results Scenario Outcomes : Environment

Results Scenario Outcomes : Environment


30
27

25

% Renewables

20

18
Reference 2003

15

Scenario A
Scenario B
Scenario C
10

10

10

Scenario D

3
1

0
2012

2025

Figure F.7: Renewables as percentage of total inland production

180,000
162,526
160,000
149,254
140,000

Heat value [TJ]

120,000

Reference 2003

100,000

Scenario A

80,935

76,821

80,000

Scenario B
Scenario C
Scenario D

60,000

40,000

35,279

37,842

37,842

37,630

37,630

20,000

0
2012

2025

Figure F.8: Total heat value of bio fuels

331

332

Appendix F Results Scenario Outcomes : Economics

Results Scenario Outcomes : Economics


6,000

5,000

4,154

Operation Cost [M]

4,000

3,689
3,287

3,298

3,334

3,242

3,317

Reference 2003

3,037
3,000

Scenario A
Scenario B
2,000

Scenario C

1,786

Scenario D

1,000

0
2012

2025

Figure F.9: Total operation cost

60

50

Operation Cost [/MWh]

40

Reference 2003
30

28

27
25

Scenario A

26

Scenario B
22

20

22

19

19

Scenario C
20

Scenario D

10

0
2012

2025

Figure F.10: Average operation cost per MWh

333

8,000

6,848

7,000

5,846

6,000

Dispatch Cost [M]

5,292
4,940

5,000

5,074

5,003
4,708
Reference 2003

4,342
4,000

Scenario A
Scenario B
Scenario C

3,000

Scenario D
2,000

1,908

1,000

0
2012

2025

Figure F.11: Total dispatch cost

60

50

42

40

Dispatch Cost [/MWh]

40

41

40

32

32

30

Reference 2003
Scenario A
Scenario B
Scenario C

20

Scenario D

20

10

0
2012

Figure F.12: Average dispatch cost per MWh

334

32

29

2025

Appendix F Results Scenario Outcomes : Economics

10,000

9,000
8,263
8,000
6,910

Dispatch + fixed cost [M]

7,000
6,294
5,938

6,000

6,180

6,075

6,088

5,779
Reference 2003

5,000

Scenario A
Scenario B
Scenario C

4,000

3,000

Scenario D
2,661

2,000

1,000

0
2012

2025

Figure F.13: Total dispact + fixed cost

60

50
50

48

49

48

Total Dispatch + Fixed Cost [/MWh]

43
41
39

40
36

Reference 2003
30

28
Scenario A
Scenario B
Scenario C
Scenario D

20

10

0
2012

2025

Figure F.14: Average dispact + fixed cost per MWh

335

12,000
11,000

10,383

Dispatch + Fixed + Financial Cost [M/a]

10,000
9,000
8,160
8,000

7,447
7,093

8,093

8,033

8,365

7,242

7,000

Reference 2003

6,000

Scenario A
Scenario B

5,000

Scenario C
Scenario D

3,825

4,000
3,000
2,000
1,000
0

2012

2025

Figure F.15: Total dispact + fixed + financial cost (= total integral cost)

70

60

Dispatch + Fixed + Financial Cost [/MWh]

60

60

58
56

57

56

49

50

47
41

40
Reference 2003
Scenario A

30

Scenario B
Scenario C
Scenario D

20

10

0
2012

2025

Figure F.16: Average dispact + fixed + financial cost per MWh (= integral cost per MWh)

336

Appendix F Results Scenario Outcomes : Economics

25,000

21,653
20,833
19,644

20,000

Investment cost [M]

17,809

15,000

Reference 2003
Scenario A
Scenario B
Scenario C

9,544

10,000

Scenario D

8,290
7,153
6,391
5,000

0
2012

2025

Figure F.17: Investement cost

1,600

1,356

1,400

1,334
1,226

1,336
1,263

Specific Investment cost [/kWe]

1,200

1,192

1,224

1,151

1,000
Reference 2003
800

Scenario A
Scenario B
Scenario C

600

Scenario D

400

200

0
2012

2025

Figure F.18: Specific investement cost

337

1,800

1,600

1,570

Cost of stimulation measures [M]

1,400

1,200

1,000

Scenario A
Scenario B

811

800

Scenario C
Scenario D

600
383
400
384
200
87
0
2000

111

87
2005

2010

2015

2020

2025

2030

Figure F.19: Cost of stimulation measures based on MEP

1,800

1,600
153

Cost of stimulation measures [M]

1,400

1,200

1,000
Bio-Energy
Renewables

800
178

1,399

CHP

600

400
621

200

53
57

Reference

18

Scenario A

17

Scenario B

Scenario C

Scenario D

Figure F.20: Breakdown of cost of stimulation measures 2025 based on MEP

338

Appendix F Results Scenario Outcomes : Security of Supply

Results Scenario Outcomes : Security of Supply


14,000
12,923

12,000

9,638

Capacity [MWe]

10,000

Reference 2003

8,000

Scenario A
Scenario B

6,000

Scenario C
Scenario D

4,542
3,692

4,000

4,097

3,377

3,900

3,451

2,000

1,503

0
2012

2025

Figure F.21: Reserve power: total installed capacity minus peak load

40
36
35

30
27

Reserve factor [%]

25
Reference 2003
20

18

Scenario A
16

Scenario B

14

15

Scenario C
Scenario D

12
10
10

10

7
5

0
2012

2025

Figure F.22: Reserve factor without import capacity and renewable capacity count for 100%

339

340

supply CCGT DH

supply "non fossil"

supply "Non Fossil"

supply "Conventional"

3,850

67

27

supply "conventional"

supply others

Import balance

414

supply "biomass incineration"

supply "Waste Incinerator

46

supply "photo-voltaic solar"

0
37

906

offshore

supply "hydro"

onshore

supply "wind"

supply "Non Fossil"

supply "non fossil"

supply "conventional"

supply Small CHP (gasengines)

1,016

528

244
1,513

797

supply CCGT DH

supply Boiler + ST

1,417

supply CCGT < 250 MWe

supply GT + HRSG

765

6,761
3,507

supply CCGT > 250 MWe

supply Industrial CHP

Supply from other electricity producers

1,826
759
1,753

17,041

236

2,041

1,626

948
4,908

20,981

7,288
4,722
9,102

176,318

17,483

26,322

398,561

0%

100%

1,996

959

1,082

205

31

72

1,318

481

4,427

2,470

5,088

7,993

5,429

146

3,484

29,139

55,033

44,141

14,369

142,682

0
12,100

610

1,184

1,890

47%

53%

10%

90%

0%

100%

0%

100%

0%

0
0
0
0

0%

100%

100%

supply CCGT CHP

supply CCGT

555,348

953,909 o.w.
35,279 Biofuel

2,797

supply "non fossil"

3,541
0
26
1,996

GR

Fuel

0%

2,802

DH

TJ

751

CHP

Heat

100%

449
0
137
253

33,135

209,453

TJ

Reference 2003

3,259

84

30,340

supply "conventional"

supply IGCC multifuel

supply "non fossil"

supply "conventional"

supply IGCC coal

supply peak

supply nuclear PBMR

supply nuclear BWR, ABWR, EPR

supply "non fossil"

supply "conventional"

2,797

3,259

30,423

12,261
18,478

30,740

63,154

17,041
110,935

63,154
30,740

106,608
4,326
110,935

110,935 o.w.
3,150 Renewable

supply "non fossil"

supply hot windbox CCGT

2,308

Electricity
GWh

2,100

13,545
3,960

12,091
4,553
3,014
18,803

20,306

MWe

Capacity

supply "conventional"

supply conventional multi fuel

supply "non fossil"

supply "conventional"

supply conventional gas

supply "non fossil"

supply "conventional"

supply conventional coal

Net supply electricity


of which:
Supply from electricity production companies

Import balance
Available domestically

CHP and other supply to public grid

own decentralized production

Supply from electricity production companies


Supply from other electricity producers

Electricity consumption
Grid losses
Total required

Highest load electricity production companies


Highest load other electricity producers
Highest load import / export
Highest load domestically

Demand - Supply

Energy balance-sheet

610

53,315

4,059

8,181
55,054

44,150

90,418

96,206

46,567

277,342

51,042
42,522
74,137

39,466
0
367
16,763

22,396

29,649

279,007

610

25,058

28,257

4,059

5,395

49,659

8,181

277,342

74,137

42,522

51,042

16,763

367

39,466

22,396

29,649

766

278,241

CO2

21,182

43,362

43

1,749

459
2,781

16,151

2,858
3,644
4,152

0
0
21
1,576

1,254

3,726

26,131

64,544 Long
4,510 Short
0 CCS

kTon

3,389

1,749

445

591

2,781

1,576

1,254

3,726

84

26,131

Appendix F Results Scenario Outcomes : Energy balance-sheet

341

342

- public grid

/GJ
/GJ
/GJ

Biomass incineration
Pure biomass (oil)
Pure biomass
Mixed biomass
> 50 MW (in /GJ)
> 50 MW (in /GJ)
> 50 MW (in /GJ)

GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

WI
Waste incineration <= 22%
Waste incineration > 22% <= 23%

> 23% <= 24%

> 24% <= 25%

> 25% <= 26%

> 26% <= 27%

> 27% <= 28%

> 28% <= 29%

> 29% <= 30%

> 30%

- public grid

CHP gas engines


- decentralized generation

- public grid

Boiler + ST
- decentralized generation

- public grid

GT + HRSG
- decentralized generation

- public grid

CHP < 120 MWe


CCGT
- decentralized generation

GWh
GWh

Unit

Stimulation Measures

CHP > 120 MWe (250 MWe)


- decentralized generation

/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/Ton

Unit

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST OTHER ELECTRICITY PRODUCERS

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST ELECTRICITY PRODUCTION COMPANIES

COAL
GAS
URANIUM
L.O
Bio-oil
Biomass Wood Pellets
Biomass Waste Wood, Agro-residues, Chicken Dung
CO2

FUEL PRICES

Economics

0.00

0.00

0.00

1,813

3,095

613

1,857

1,005

4,084

5,103

3,226

5,429

Value

0
176,956,098
0
0
98,150
8,165,200
4,483,877

0
1,351,173,235
0
0
23,649,980
97,768,289
177,814,892

1.50
3.80
1.00
10.00
9.50
6.00
3.50
0.00

Value

UNIT-34
UNIT-35
UNIT-36
UNIT-37
UNIT-38
UNIT-39
UNIT-40
UNIT-41
UNIT-42
UNIT-43
UNIT-44
UNIT-45
UNIT-46
UNIT-47
UNIT-48
UNIT-49
UNIT-50
UNIT-51
UNIT-52
UNIT-53
UNIT-54
UNIT-55
PAC

UNIT-1
UNIT-2
UNIT-3
UNIT-4
UNIT-5
UNIT-6
UNIT-7
UNIT-8
UNIT-9
UNIT-10
UNIT-11
UNIT-12
UNIT-13
UNIT-14
UNIT-15
UNIT-16
UNIT-17
UNIT-18
UNIT-19
UNIT-20
UNIT-21
UNIT-22
UNIT-23
UNIT-24
UNIT-25
UNIT-26
UNIT-27
UNIT-28
UNIT-29
UNIT-30
UNIT-31
UNIT-32
UNIT-33

Naam

Type

PAC

0
NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

600
600
1200
1200
1200
500
500
500
500
1000
1000
1465
1465
1600
1600
300
300
300
300
300
0
0
1500

Sel Unit Size


MWe
0
1200
0
1200
0
1200
0
1200
0
550
0
550
0
550
0
550
0
900
0
900
0
480
0
480
0
480
0
480
0
480
0
480
0
832
0
832
0
832
0
170
0
170
0
170
0
170
0
1200
0
1200
0
1200
0
1200
0
1200
0
1200
0
600
0
600
0
600
0
600

IGCC

IGCC
IGCC

IGCC
IGCC

CONV MULT

CONV MULT

CONV COAL

CONV COAL

CONV COAL

CONV COAL

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

IGCC

IGCC

CONV MULT

CONV COAL

CONV COAL

CONV COAL

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

MWe

Y
N
N
Y
Y
N
N
N
N
N
N
-

Y/N
N
N
N
N
Y
Y
Y
Y
Y
Y

Installed CCS

525

supply Boiler + ST

supply "biomass incineration"

supply "photo-voltaic solar"

supply "Non Fossil"

supply "Conventional"

4,450

67
2,438

236

39,031

673,448

0%

100%

205

1,011

72

18,690

4,501

717

6,604

2,470

11,492

16,500

4,328

6,930

13,010

1,064

146

3,484

14,345

258,892

0
19,656

6,603

1,184

1,890

47%

53%

10%

90%

35%

65%

0%

100%

0%

0
0
0
0

20%

80%

100%

2,675

supply others

15,972

540,403

1,213,851 o.w.
149,254 Biofuel

supply "non fossil"

5,692

311,229

GR

Fuel

0%

8,538

DH

TJ

711

CHP

Heat

100%

3,017

27

1,500

24,479

948
7,321

34,790

6,753
4,620
7,100

19,940

3,475
0
2
1,064

44,878

356,107

TJ

Scenario A_2025

2,349

7,013

supply "conventional"

supply "Waste Incinerator

Import balance

1,150

6,000
37

2,000

offshore

supply "hydro"

onshore

supply "wind"

supply "Non Fossil"

supply "non fossil"

supply "conventional"

supply Small CHP (gasengines)

9,564

1,800

supply GT + HRSG

244
2,250

2,925

supply CCGT < 250 MWe

supply CCGT DH

765

19,290
6,015

6,475
835
1,479

3,300

449
0
120
253

supply CCGT > 250 MWe

supply Industrial CHP

Supply from other electricity producers

supply CCGT DH

supply CCGT CHP

supply CCGT

supply "non fossil"

supply "conventional"

supply IGCC multifuel

supply "non fossil"

supply "conventional"

supply IGCC coal

supply peak

supply nuclear PBMR

supply nuclear BWR, ABWR, EPR

supply "non fossil"

supply "conventional"

9,878

2,349

7,013

22,934
50,533

1,976

73,467

62,193

2,438
138,099

62,193
73,467

132,713
5,386
138,099

138,099 o.w.
36,777 Renewable

supply "non fossil"

supply hot windbox CCGT

Electricity
GWh

7,903

2,400

361

16,902
1,230

15,477
13,519
600
23,269

36,192

MWe

Capacity

supply "conventional"

supply conventional multi fuel

supply "non fossil"

supply "conventional"

supply conventional gas

supply "non fossil"

supply "conventional"

supply conventional coal

Net supply electricity


of which:
Supply from electricity production companies

Import balance
Available domestically

CHP and other supply to public grid

own decentralized production

Supply from electricity production companies


Supply from other electricity producers

Electricity consumption
Grid losses
Total required

Highest load electricity production companies


Highest load other electricity producers
Highest load import / export
Highest load domestically

Demand - Supply

Energy balance-sheet

610

133,206

4,059

8,181
75,024

40,351

452,368

38,603
49,675
62,451

166,553

38,761
0
25
9,022

86,768

21,369

67,175

610

62,607

70,599

4,059

7,352

67,672

8,181

452,368

62,451

49,675

38,603

57,883

108,670

9,022

25

38,761

17,354

69,415

21,369

67,175

CO2

34,824

22,333

2,162
4,037
3,497

-391

0
0
1
848

2,755

3,205

6,219

43

4,370

459
3,790

26,163

57,158 Long
17,967 Short
14,386 CCS

kTon

8,468

4,370

445

806

3,790

6,346

-391

848

1,902

2,755

3,205

6,219

Appendix F Results Scenario Outcomes : Energy balance-sheet

343

344

- public grid

/GJ
/GJ
/GJ

Biomass incineration
Pure biomass (oil)
Pure biomass
Mixed biomass
> 50 MW (in /GJ)
> 50 MW (in /GJ)
> 50 MW (in /GJ)

GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

WI
Waste incineration <= 22%
Waste incineration > 22% <= 23%

> 23% <= 24%

> 24% <= 25%

> 25% <= 26%

> 26% <= 27%

> 27% <= 28%

> 28% <= 29%

> 29% <= 30%

> 30%

- public grid

CHP gas engines


- decentralized generation

- public grid

Boiler + ST
- decentralized generation

- public grid

GT + HRSG
- decentralized generation

- public grid

CHP < 120 MWe


CCGT
- decentralized generation

GWh
GWh

Unit

Stimulation Measures

CHP > 120 MWe (250 MWe)


- decentralized generation

/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/Ton

Unit

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST OTHER ELECTRICITY PRODUCERS

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST ELECTRICITY PRODUCTION COMPANIES

COAL
GAS
URANIUM
L.O
Bio-oil
Biomass Wood Pellets
Biomass Waste Wood, Agro-residues, Chicken Dung
CO2

FUEL PRICES

Economics

0.00

0.00

0.00

515

257

906

619

995

2,115

5,207

624

1,846

2,269

9,223

10,699

6,658

4,328

Value

0
280,485,573
0
44,877,689
114,510
6,509,310
9,441,139

0
2,043,572,189
0
554,660,512
39,277,247
117,104,262
183,542,221

2.00
7.00
1.20
17.50
9.50
6.00
3.50
20.00

Value

UNIT-34
UNIT-35
UNIT-36
UNIT-37
UNIT-38
UNIT-39
UNIT-40
UNIT-41
UNIT-42
UNIT-43
UNIT-44
UNIT-45
UNIT-46
UNIT-47
UNIT-48
UNIT-49
UNIT-50
UNIT-51
UNIT-52
UNIT-53
UNIT-54
UNIT-55
PAC

UNIT-1
UNIT-2
UNIT-3
UNIT-4
UNIT-5
UNIT-6
UNIT-7
UNIT-8
UNIT-9
UNIT-10
UNIT-11
UNIT-12
UNIT-13
UNIT-14
UNIT-15
UNIT-16
UNIT-17
UNIT-18
UNIT-19
UNIT-20
UNIT-21
UNIT-22
UNIT-23
UNIT-24
UNIT-25
UNIT-26
UNIT-27
UNIT-28
UNIT-29
UNIT-30
UNIT-31
UNIT-32
UNIT-33

Naam

Type

PAC

0
NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

600
600
1200
1200
1200
500
500
500
500
1000
1000
1465
1465
1600
1600
300
300
300
300
300
0
0
1500

Sel Unit Size


MWe
0
1200
0
1200
0
1200
1
1200
0
550
0
550
0
550
0
550
1
900
0
900
0
480
0
480
0
480
0
480
0
480
0
480
0
832
0
832
0
832
0
170
0
170
0
170
0
170
0
1200
0
1200
0
1200
0
1200
1
1200
0
1200
0
600
0
600
0
600
0
600

IGCC

IGCC
IGCC

IGCC
IGCC

CONV MULT

CONV MULT

CONV COAL

CONV COAL

CONV COAL

CONV COAL

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

IGCC

IGCC

CONV MULT

CONV COAL

CONV COAL

CONV COAL

0
0
0
1200
1200
500
500
500
500
1000
1000
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
1200
0
0
0
0
900
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1200
0
0
0
0
0

MWe

Y
N
N
Y
Y
N
N
N
N
N
N
-

Y/N
N
N
N
N
Y
Y
Y
Y
Y
Y

Installed CCS

supply "non fossil"

supply CCGT DH

28,712

supply "Non Fossil"

supply "Conventional"

3,850

1,131

236

205

supply "non fossil"

2,407

31

72

481

4,427

2,470

5,088

7,993

4,354

1,275

67

27

308

948
4,908

19,906

2,468
4,641
7,102

36,921

supply "conventional"

supply others

Import balance

485

supply "biomass incineration"

supply "Waste Incinerator

46

supply "photo-voltaic solar"

0
37

offshore

supply "hydro"

onshore

supply "wind"

supply "Non Fossil"

supply "non fossil"

supply "conventional"

supply Small CHP (gasengines)

110

528

244
1,513

797

supply CCGT DH

supply Boiler + ST

1,417

supply CCGT < 250 MWe

supply GT + HRSG

765

5,926
3,507

supply CCGT > 250 MWe

supply Industrial CHP

Supply from other electricity producers

2,475
835
1,479
176,175

16,576

26,191

379,269

0%

146

3,484

29,139

55,033

44,141

14,357

142,670

0
19,659

610

1,184

1,890

47%

53%

10%

90%

0%

100%

0%

100%

0%

0
0
0
0

100%

supply CCGT CHP

supply CCGT

0%

100%

100%

supply "non fossil"

52,917
0
17
36,921

1,210,398

1,589,667 o.w.
37,630 Biofuel

6,579
0
120
6,053

supply "conventional"

supply IGCC multifuel

supply "non fossil"

supply "conventional"

supply IGCC coal

supply peak

supply nuclear PBMR

supply nuclear BWR, ABWR, EPR

supply "conventional"

supply hot windbox CCGT

GR

Fuel

0%

2,528

DH

TJ

711

CHP

Heat

100%

supply "non fossil"

39,474

215,649

TJ

Scenario B_2025

2,354

35,895

2,354

35,895

12,261
16,451

supply "conventional"

supply conventional multi fuel

supply "non fossil"

supply "conventional"

supply conventional gas

supply "non fossil"

supply "conventional"

361

142,314

Net supply electricity


of which:
Supply from electricity production companies

supply conventional coal

0
171,026

Import balance
Available domestically

CHP and other supply to public grid

own decentralized production

142,314
28,712

171,026 o.w.
1,920 Renewable

Supply from electricity production companies


Supply from other electricity producers

26,332
8,430

Electricity
GWh

164,356
6,670
171,026

24,815
4,068
0
28,880

32,258

MWe

Capacity

Electricity consumption
Grid losses
Total required

Highest load electricity production companies


Highest load other electricity producers
Highest load import / export
Highest load domestically

Demand - Supply

Energy balance-sheet

610

60,890

4,059

8,181
50,541

44,150

83,715

86,618

40,505

254,988

16,306
49,759
61,678

534,977
0
185
255,983

21,416

270,093

610

28,618

32,272

4,059

4,953

45,588

8,181

254,988

61,678

49,759

16,306

255,983

185

534,977

21,416

270,093

CO2

19,934

61,041

43

1,997

459
2,553

14,882

913
4,041
3,454

0
0
10
24,062

3,212

25,347

80,975 Long
4,859 Short
0 CCS

kTon

3,871

1,997

445

543

2,553

24,062

3,212

25,347

Appendix F Results Scenario Outcomes : Energy balance-sheet

345

346

- public grid

/GJ
/GJ
/GJ

Biomass incineration
Pure biomass (oil)
Pure biomass
Mixed biomass
> 50 MW (in /GJ)
> 50 MW (in /GJ)
> 50 MW (in /GJ)

GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

WI
Waste incineration <= 22%
Waste incineration > 22% <= 23%

> 23% <= 24%

> 24% <= 25%

> 25% <= 26%

> 26% <= 27%

> 27% <= 28%

> 28% <= 29%

> 29% <= 30%

> 30%

- public grid

CHP gas engines


- decentralized generation

- public grid

Boiler + ST
- decentralized generation

- public grid

GT + HRSG
- decentralized generation

- public grid

CHP < 120 MWe


CCGT
- decentralized generation

GWh
GWh

Unit

Stimulation Measures

CHP > 120 MWe (250 MWe)


- decentralized generation

/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/Ton

Unit

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST OTHER ELECTRICITY PRODUCERS

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST ELECTRICITY PRODUCTION COMPANIES

COAL
GAS
URANIUM
L.O
Bio-oil
Biomass Wood Pellets
Biomass Waste Wood, Agro-residues, Chicken Dung
CO2

FUEL PRICES

Economics

0.00

0.00

0.00

1,813

3,095

613

1,857

1,005

4,084

5,635

3,226

4,354

Value

0
283,532,153
0
45,365,130
114,510
6,548,050
9,298,006

0
2,711,926,738
0
1,218,457,951
36,993,485
158,142,188
105,745,152

2.00
7.00
1.20
17.50
9.50
6.00
3.50
20.00

Value

UNIT-34
UNIT-35
UNIT-36
UNIT-37
UNIT-38
UNIT-39
UNIT-40
UNIT-41
UNIT-42
UNIT-43
UNIT-44
UNIT-45
UNIT-46
UNIT-47
UNIT-48
UNIT-49
UNIT-50
UNIT-51
UNIT-52
UNIT-53
UNIT-54
UNIT-55
PAC

UNIT-1
UNIT-2
UNIT-3
UNIT-4
UNIT-5
UNIT-6
UNIT-7
UNIT-8
UNIT-9
UNIT-10
UNIT-11
UNIT-12
UNIT-13
UNIT-14
UNIT-15
UNIT-16
UNIT-17
UNIT-18
UNIT-19
UNIT-20
UNIT-21
UNIT-22
UNIT-23
UNIT-24
UNIT-25
UNIT-26
UNIT-27
UNIT-28
UNIT-29
UNIT-30
UNIT-31
UNIT-32
UNIT-33

Naam

Type

PAC

0
NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

600
600
1200
1200
1200
500
500
500
500
1000
1000
1465
1465
1600
1600
300
300
300
300
300
0
0
1500

Sel Unit Size


MWe
1
1200
1
1200
1
1200
0
1200
1
550
1
550
1
550
1
550
0
900
0
900
0
480
0
480
0
480
0
480
0
480
0
480
0
832
0
832
0
832
0
170
0
170
0
170
0
170
1
1200
1
1200
1
1200
0
1200
0
1200
0
1200
1
600
1
600
1
600
1
600

IGCC

IGCC
IGCC

IGCC
IGCC

CONV MULT

CONV MULT

CONV COAL

CONV COAL

CONV COAL

CONV COAL

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

IGCC

IGCC

CONV MULT

CONV COAL

CONV COAL

CONV COAL

0
0
1200
0
0
0
0
0
0
0
0
1465
1465
1600
1600
0
0
0
0
0
0
0
0

1200
1200
1200
0
550
550
550
550
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1200
1200
1200
0
0
0
600
600
600
600

MWe

N
N
N
N
N
N
N
N
N
N
N
-

Y/N
N
N
N
N
N
N
N
N
N
N

Installed CCS

525

supply Boiler + ST

supply "biomass incineration"

supply "photo-voltaic solar"

supply "Non Fossil"

supply "Conventional"

5,050

4,877

15,975

34,684

GR

581,623

806,204

0%

100%

1,387,827 o.w.
162,526 Biofuel

Fuel

205

505

72

9,495

3,376

637

5,865

2,470

8,300

13,679

4,248

7,638

14,388

21,378

146

3,484

29,139

89,765

75,542

14,345

208,791

0
19,657

6,603

1,184

1,890

47%

53%

10%

90%

35%

65%

0%

100%

0%

0
0
0
0

20%
100%

0
0

2,791

80%

0%

11,164

8,558

DH

TJ

711

CHP

Heat

100%

256,782

44,901

301,683

TJ

Scenario C_2025

2,345

13,321

2,675

236

5,692

13,653

948
6,503

28,698

4,665
4,597
6,968

22,026

3,780
0
1
21,378

13,954

2,345

13,321

18,479
37,251

supply "non fossil"

supply others

67

55,730

93,036

4,877
153,643

93,036
55,730

147,651
5,992
153,643

153,643 o.w.
27,395 Renewable

3,017

27

750

Electricity
GWh

supply "conventional"

supply "Waste Incinerator

Import balance

1,150

3,000
37

1,500

offshore

supply "hydro"

onshore

supply "wind"

supply "Non Fossil"

supply "non fossil"

supply "conventional"

supply Small CHP (gasengines)

5,314

1,300

supply GT + HRSG

244
2,000

2,425

supply CCGT < 250 MWe

supply CCGT DH

765

13,790
5,015

5,975
835
1,479

3,300

449
0
120
3,253

3,600

361

21,802
2,430

18,624
9,752
1,200
25,953

35,592

MWe

Capacity

supply CCGT > 250 MWe

supply Industrial CHP

Supply from other electricity producers

supply CCGT DH

supply CCGT CHP

supply CCGT

supply "non fossil"

supply "conventional"

supply IGCC multifuel

supply "non fossil"

supply "conventional"

supply IGCC coal

supply peak

supply nuclear PBMR

supply nuclear BWR, ABWR, EPR

supply "non fossil"

supply "conventional"

supply hot windbox CCGT

supply "non fossil"

supply "conventional"

supply conventional multi fuel

supply "non fossil"

supply "conventional"

supply conventional gas

supply "non fossil"

supply "conventional"

supply conventional coal

Net supply electricity


of which:
Supply from electricity production companies

Import balance
Available domestically

CHP and other supply to public grid

own decentralized production

Supply from electricity production companies


Supply from other electricity producers

Electricity consumption
Grid losses
Total required

Highest load electricity production companies


Highest load other electricity producers
Highest load import / export
Highest load domestically

Demand - Supply

Energy balance-sheet

610

133,206

4,059

8,181
66,733

44,150

136,548

148,236

39,901

368,835

26,479
49,452
61,480

183,374

42,001
0
15
175,775

128,642

21,332

117,653

610

62,607

70,599

4,059

6,540

60,193

8,181

368,835

61,480

49,452

26,479

63,592

119,781

175,775

15

42,001

25,728

102,914

21,332

117,653

CO2

29,624

27,021

43

4,370

459
3,371

21,382

1,483
4,024
3,443

-359

0
0
1
2,717

1,528

3,200

10,984

56,645 Long
19,422 Short
33,602 CCS

kTon

8,468

4,370

445

717

3,371

6,972

-359

2,717

2,821

1,528

3,200

10,984

Appendix F Results Scenario Outcomes : Energy balance-sheet

347

348

- public grid

/GJ
/GJ
/GJ

Biomass incineration
Pure biomass (oil)
Pure biomass
Mixed biomass
> 50 MW (in /GJ)
> 50 MW (in /GJ)
> 50 MW (in /GJ)

GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

WI
Waste incineration <= 22%
Waste incineration > 22% <= 23%

> 23% <= 24%

> 24% <= 25%

> 25% <= 26%

> 26% <= 27%

> 27% <= 28%

> 28% <= 29%

> 29% <= 30%

> 30%

- public grid

CHP gas engines


- decentralized generation

- public grid

Boiler + ST
- decentralized generation

- public grid

GT + HRSG
- decentralized generation

- public grid

CHP < 120 MWe


CCGT
- decentralized generation

GWh
GWh

Unit

Stimulation Measures

CHP > 120 MWe (250 MWe)


- decentralized generation

/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/Ton

Unit

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST OTHER ELECTRICITY PRODUCERS

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST ELECTRICITY PRODUCTION COMPANIES

COAL
GAS
URANIUM
L.O
Bio-oil
Biomass Wood Pellets
Biomass Waste Wood, Agro-residues, Chicken Dung
CO2

FUEL PRICES

Economics

0.00

0.00

0.00

515

257

906

619

995

2,051

4,452

624

1,846

1,639

6,661

9,035

5,520

4,248

Value

0
277,489,418
0
44,398,291
114,510
6,389,510
9,869,091

0
2,544,525,452
0
676,725,862
33,071,497
190,252,941
178,894,738

2.00
7.00
1.20
17.50
9.50
6.00
3.50
20.00

Value

UNIT-34
UNIT-35
UNIT-36
UNIT-37
UNIT-38
UNIT-39
UNIT-40
UNIT-41
UNIT-42
UNIT-43
UNIT-44
UNIT-45
UNIT-46
UNIT-47
UNIT-48
UNIT-49
UNIT-50
UNIT-51
UNIT-52
UNIT-53
UNIT-54
UNIT-55
PAC

UNIT-1
UNIT-2
UNIT-3
UNIT-4
UNIT-5
UNIT-6
UNIT-7
UNIT-8
UNIT-9
UNIT-10
UNIT-11
UNIT-12
UNIT-13
UNIT-14
UNIT-15
UNIT-16
UNIT-17
UNIT-18
UNIT-19
UNIT-20
UNIT-21
UNIT-22
UNIT-23
UNIT-24
UNIT-25
UNIT-26
UNIT-27
UNIT-28
UNIT-29
UNIT-30
UNIT-31
UNIT-32
UNIT-33

Naam

Type

PAC

0
NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

600
600
1200
1200
1200
500
500
500
500
1000
1000
1465
1465
1600
1600
300
300
300
300
300
0
0
1500

Sel Unit Size


MWe
1
1200
0
1200
0
1200
1
1200
0
550
0
550
0
550
0
550
1
900
0
900
0
480
0
480
0
480
0
480
0
480
0
480
0
832
0
832
0
832
0
170
0
170
0
170
0
170
0
1200
0
1200
0
1200
0
1200
1
1200
1
1200
1
600
1
600
1
600
1
600

IGCC

IGCC
IGCC

IGCC
IGCC

CONV MULT

CONV MULT

CONV COAL

CONV COAL

CONV COAL

CONV COAL

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

IGCC

IGCC

CONV MULT

CONV COAL

CONV COAL

CONV COAL

600
0
0
1200
1200
500
500
500
0
1000
1000
0
0
0
0
0
0
0
0
0
0
0
0

1200
0
0
1200
0
0
0
0
900
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1200
1200
600
600
600
600

MWe

Y
N
N
Y
Y
N
N
N
N
N
N
-

Y/N
N
N
N
N
Y
Y
Y
Y
Y
Y

Installed CCS

supply CCGT DH

28,828

52,823

supply "Non Fossil"

supply "Conventional"

3,850

3,850

1,131

236

205

31

72

481

4,427

2,470

5,088

7,993

4,469

supply "non fossil"

2,407

176,187

16,564

26,191

379,912

0%

146

3,484

29,139

55,033

44,141

14,369

142,682

0
19,657

610

1,184

1,890

47%

53%

10%

90%

0%

100%

0%

100%

0%

0
0
0
0

100%

1,275

67

27

308

948
4,908

20,021

12,822
4,711
7,633

0%

100%

100%

supply "conventional"

supply others

Import balance

485

supply "biomass incineration"

supply "Waste Incinerator

46

supply "photo-voltaic solar"

0
37

offshore

supply "hydro"

onshore

supply "wind"

supply "Non Fossil"

supply "non fossil"

supply "conventional"

supply Small CHP (gasengines)

110

528

244
1,513

797

supply CCGT DH

supply Boiler + ST

1,417

supply CCGT < 250 MWe

supply GT + HRSG

765

5,926
3,507

supply CCGT > 250 MWe

supply Industrial CHP

Supply from other electricity producers

8,059
835
1,543

1,406,852

1,786,763 o.w.
37,630 Biofuel

supply CCGT CHP

supply CCGT

GR

Fuel

0%

17,024

2,534

DH

TJ

711

CHP

Heat

100%

supply "non fossil"

28,339
0
8
52,823

17,024

39,465

215,652

TJ

Scenario D_2025

2,353

57,437

3,514
0
120
7,253

2,400

2,353

57,437

12,261
16,567

supply "conventional"

supply IGCC multifuel

supply "non fossil"

supply "conventional"

supply IGCC coal

supply peak

supply nuclear PBMR

supply nuclear BWR, ABWR, EPR

supply "non fossil"

supply "conventional"

supply hot windbox CCGT

supply "non fossil"

supply "conventional"

supply conventional multi fuel

supply "non fossil"

supply "conventional"

supply conventional gas

supply "non fossil"

supply "conventional"

361

183,149

Net supply electricity


of which:
Supply from electricity production companies

supply conventional coal

0
211,977

Import balance
Available domestically

CHP and other supply to public grid

own decentralized production

183,149
28,828

211,977 o.w.
1,920 Renewable

Supply from electricity production companies


Supply from other electricity producers

33,715
9,630

Electricity
GWh

203,710
8,267
211,977

31,675
4,068
0
35,740

39,641

MWe

Capacity

Electricity consumption
Grid losses
Total required

Highest load electricity production companies


Highest load other electricity producers
Highest load import / export
Highest load domestically

Demand - Supply

Energy balance-sheet

610

60,890

4,059

8,181
50,541

44,150

83,715

86,618

41,147

255,630

75,834
50,052
65,744

288,482
0
95
363,178

116,741

21,409

425,317

610

28,618

32,272

4,059

4,953

45,588

8,181

255,630

65,744

50,052

75,834

363,178

95

288,482

116,741

21,409

425,317

CO2

19,970

100,259

43

1,997

459
2,553

14,918

4,247
4,061
3,682

0
0
5
34,139

10,974

3,211

39,940

120,229 Long
4,859 Short
0 CCS

kTon

3,871

1,997

445

543

2,553

34,139

10,974

3,211

39,940

Appendix F Results Scenario Outcomes : Energy balance-sheet

349

350

- public grid

/GJ
/GJ
/GJ

Biomass incineration
Pure biomass (oil)
Pure biomass
Mixed biomass
> 50 MW (in /GJ)
> 50 MW (in /GJ)
> 50 MW (in /GJ)

GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

GWh
GWh

WI
Waste incineration <= 22%
Waste incineration > 22% <= 23%

> 23% <= 24%

> 24% <= 25%

> 25% <= 26%

> 26% <= 27%

> 27% <= 28%

> 28% <= 29%

> 29% <= 30%

> 30%

- public grid

CHP gas engines


- decentralized generation

- public grid

Boiler + ST
- decentralized generation

- public grid

GT + HRSG
- decentralized generation

- public grid

CHP < 120 MWe


CCGT
- decentralized generation

GWh
GWh

Unit

Stimulation Measures

CHP > 120 MWe (250 MWe)


- decentralized generation

/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/GJ
/Ton

Unit

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST OTHER ELECTRICITY PRODUCERS

FIXED
FUEL
NOX
CO2
START
VOM
CAP

COST ELECTRICITY PRODUCTION COMPANIES

COAL
GAS
URANIUM
L.O
Bio-oil
Biomass Wood Pellets
Biomass Waste Wood, Agro-residues, Chicken Dung
CO2

FUEL PRICES

Economics

0.00

0.00

0.00

1,813

3,095

613

1,857

1,005

4,084

5,677

3,226

4,469

Value

0
288,029,705
0
46,084,770
114,510
6,720,620
8,720,093

0
3,619,292,523
0
2,000,630,883
68,557,859
277,553,919
208,064,690

2.00
7.00
1.20
17.50
9.50
6.00
3.50
20.00

Value

UNIT-34
UNIT-35
UNIT-36
UNIT-37
UNIT-38
UNIT-39
UNIT-40
UNIT-41
UNIT-42
UNIT-43
UNIT-44
UNIT-45
UNIT-46
UNIT-47
UNIT-48
UNIT-49
UNIT-50
UNIT-51
UNIT-52
UNIT-53
UNIT-54
UNIT-55
PAC

UNIT-1
UNIT-2
UNIT-3
UNIT-4
UNIT-5
UNIT-6
UNIT-7
UNIT-8
UNIT-9
UNIT-10
UNIT-11
UNIT-12
UNIT-13
UNIT-14
UNIT-15
UNIT-16
UNIT-17
UNIT-18
UNIT-19
UNIT-20
UNIT-21
UNIT-22
UNIT-23
UNIT-24
UNIT-25
UNIT-26
UNIT-27
UNIT-28
UNIT-29
UNIT-30
UNIT-31
UNIT-32
UNIT-33

Naam

Type

PAC

0
NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

600
600
1200
1200
1200
500
500
500
500
1000
1000
1465
1465
1600
1600
300
300
300
300
300
0
0
1500

Sel Unit Size


MWe
1
1200
1
1200
1
1200
0
1200
1
550
1
550
1
550
1
550
0
900
0
900
1
480
1
480
1
480
1
480
0
480
0
480
1
832
1
832
0
832
0
170
0
170
0
170
0
170
1
1200
1
1200
1
1200
1
1200
1
1200
1
1200
1
600
1
600
1
600
1
600

IGCC

IGCC
IGCC

IGCC
IGCC

CONV MULT

CONV MULT

CONV COAL

CONV COAL

CONV COAL

CONV COAL

NUCLEAR

NUCLEAR

NUCLEAR

NUCLEAR

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

CCGT

IGCC MULTI

IGCC MULTI

IGCC

IGCC

IGCC

IGCC

CONV MULT

CONV COAL

CONV COAL

CONV COAL

600
600
1200
0
0
500
500
500
500
0
0
1465
0
1600
0
0
0
0
0
0
0
0
0

1200
1200
1200
0
550
550
550
550
0
0
480
480
480
480
0
0
832
832
0
0
0
0
0
1200
1200
1200
1200
1200
1200
600
600
600
600

MWe

N
N
N
N
N
N
N
N
N
N
N
-

Y/N
N
N
N
N
N
N
N
N
N
N

Installed CCS

Nomenclature

Nomenclature

abbreviations, acronyms and subscripts:


ABFB
ABWR
AC
AEB
AER
ALWR
AOO
APX
AVI
BBL
BFBC
BFG
BMC
BWR
CAES
CBS
CC
CCGT
CCT
CFBC
CHP
COCONUT
COVRA
CPB
CSP
DC
DC
DG
DH
DSM
DTE
E
EB
EC
ECN
EIA
EIA
ENERGIENED
EOH
EPC
EPON
EPR
EPRI
ESP
EU
EZ
FBC
FEGT
FGD
GDP
GE
GE
GKN

= Atmospheric Bubbling Fluidized Bed


= Advanced Boiling Water Reactor
= Alternating Current
= Afval Energie Bedrijf
= Algemene Energieraad (Dutch Energy Council)
= Advanced Light Water Reactors
= Afval Overleg Orgaan (Waste consultation body)
= Amsterdam Power Exchange
= Afval Verbrandings Installatie (Waste Incinerator)
= Balgzand Bacton Line
= Bubbling Fluidized Bed Combustion
= Blast Furnace Gas
= Biomassacentrale (Biomass plant)
= Boiling Water Reactor
= Compresses Air Energy Storage
= Centraal Bureau voor Statistiek (Central bureau of statistics)
= Combined Cycle
= Combined Cycle Gas Turbine
= Clean Coal Technologies
= Circulating Fluidized Bed Combustion
= Combined Heat and Power
= Commissie Concentratie van Nutsbedrijven (Concentration of Utilities Committee)
= Central Organization for Radioactive Waste Borssele
= Centraal Planbureau (Netherlands Bureau for Economic Policy Analysis)
= Concentrating Solar Power
= Direct Current
= District Cooling
= Distributed Generation
= District Heating
= Demand Side Management
= Dienst Toezicht Energie (Office of Energy Regulation)
= Electricity
= Energiebelasting (Energy tax)
= European Commission
= Energieonderzoek Centrum Nederland (Energy research Centre of the Netherlands)
= Energie Investerings Aftrek (Energy investment allowance)
= Energy Information Administration
= Vereniging van Energiedistributiebedrijven in Nederland (Association of
Energy Companies in the Netherlands)
= Equivalent Operating Hours
= Engineering, Procurement and Construction
= Electriciteits-Productiemaatschappij Oost- en Noord-Nederland
= European Pressurized Reactor
= Electric Power Research Institute
= Electrostatic precipitators
= European Union
= Ministerie van Economische Zaken (Ministry of Economic Affairs)
= Fluidized Bed Combustion
= Flue Exit Gas Temperature
= Flue Gas Desulfurization
= Gross Domestic Product
= Global Economy
= General Electric
= Gemeenschappelijke Kernenergiecentrale Nederland

351

GTS
GWP
H
HRSG
HTR
HVC
HVDC
ICAD
IEA
IGCC
IPCC
IPT
ITER
KAN
KEMA
LAP
LBT
LEO
LHV
LNB
LOLE
LVTE
MAP
MBDOE
MEP
MER
MNP
MT
MTS
NA
NBR
NGO
NMDA
NRG
OECD
OPAC
ORC
OSA
OVS
PAC
PBMR
PCC
PCFBC
PCU
PFBC
PGEM
PR
PTU
PWR
PV
RBR
RBT
RCCV
RCIC
RIVM
RNB
ROI

352

= Gas Transport Services


= Global warming potentials
= Heat
= Heat Recovery Steam Generator
= High Temperature Reactor
= Huisvuilcentrale
= High Voltage Direct Current
= Intercooled Aero-Derivative
= International Energy Agency
= Integrated gasification combined cycle
= Intergovernmental Panel Climate Change
= Integral Pipe Tunnel
= International Tokamak Experimental Reactor
= Knooppunt Arnhem Nijmegen (Region comprising the Dutch cities of Arnhem and
Nijmegen)
= N.V. tot Keuring van Elektrotechnische Materialen (the Electrical Engineering
Equipment Testing Company)
= Landelijk Afval Plan (National Waste Plan)
= Landelijk Basis Tarief (National Base Tariff)
= Landelijke Economische Optimalisatie (National economic optimization)
= Lower Heating Value
= Landelijk netbeheer (National gridcontrol)
= Loss of Load Expectation
= Lange Termijn Visie Energievoorziening (Long-term vision of energy supply)
= Milieu Actie Plan (Environmental action plan)
= Millions of Barrels per Day of Oil Equivalent
= Milieukwaliteit elektriciteitsproductie (Subsidy scheme for electricity)
= Milieu effect rapportage (Environmental effect report)
= Milieu en Natuur Planbureau (Netherlands Environment Assessment Agency)
= Metric Ton
= Main Transport System
= Not Applicable
= National Basic Rate (see also LBT)
= Non Governmental Organisation
= Niet Meer Dan Anders (Not more then else)
= Nuclear Regulatory Commission
= Organisation for Economic Co-operation and Development
= Underground pumped storage hydropower
= Organic Rankine Cycle
= Operation Simulation Associates
= Overeenkomst van samenwerking (Agreement of cooperation)
= Pumped Storage Hydropower
= Pebble Bed Modular Reactor
= Pulverised Coal Combustion
= Pressurized Circulating Fluidized Bed Combustion
= Power Conversion Unit
= Pressurized Fluidized Bed Combustion
= Provinciale Gelderse Energie-Maatschappij
= Program Responsible
= Program Time Unit
= Pressurized Water Reactor
= Photo Voltaic
= Regional basic rate (see also RBT)
= Regionaal basistarief
= Reinforced Concrete Containment Vessel
= Reactor Core Isolation Cooling
= Rijksinstituut voor Volksgezondheid en Milieu (National Institute for Public Health
and the Environment)
= Regionaal netbeheer (Regional gridcontrol)
= Return On Investment

Nomenclature

RPV
RTS
SC
SCR
SDE
SEP
SG
SMES
SNRC
Th
TOE
TPA
TSO
UCTE
UK
VAMIL
VAT
VDEN
VEEN
VROM
WACC
WEC
WED
WI

= Reactor Pressure Vessel


= Regional Transport System
= Single Cycle
= Selective Catalytic Reduction
= Stimuleringsregeling duurzame energieproductie (stimulation regulation
renewable energy production)
= N.V. Samenwerkende Elektriciteits-Productiebedrijven (Association of Energy
Producers)
= Steam Generator
= Super-conducting magnetic energy storage
= Selective Non-catalytic reduction
= Thermal
= Tonne of Oil Equivalent
= Third Party Access
= Transmission System Operator
= Union for the Co-ordination of Transmission of Electricity
= Uitvoeringsnota Klimaatbeleid
= Vervroegde Afschrijving Milieu Investeringen (Earlier depreciation environmental
investments)
= Value Added Tax
= Vereniging van Directeuren van Elektriciteitsbedrijven in Nederland (Associated
electricity company managers)
= Vereniging van Exploitanten van Elektriciteitsbedrijven in Nederland (Associated
Electicity Company Operators)
= Ministerie van Volkshuisvesting, Ruimtelijke Ordening en Milieubeheer (Netherlands
Ministery of Housing, Spatial Planning and the Environment)
= Weighted Average Cost of Capital
= World Energy Council
= World Energy Dialogue
= Waste Incineration

components
CH4
CO
CO2
COS
H2S
HCN
HDPE
LNG
MEA
NH3
N2O
NOx
PE
PEX
PSB
PUR
SO2

= Methane
= Carbon monoxide
= Carbon dioxide
= Carbonyl sulfide
= Hydrogen sulfide
= Hydrogen cyanide
= High density polyethylene
= Liquefied natural gas
= Mono ethanol amine
= Ammonia
= Nitrogen dioxides
= Nitrogen oxides
= Polyethylene
= Cross-linked polyethylene
= Poly sulphide bromide
= Polyurethane
= Sulpher dioxide

units
MWth
MWe
MVA
EJ
PJ

= Megawatt thermal power representing the fuel input or heat available


= Megawatt electric power output
= Mega Volt Ampere
= 1,000 PJ, energy value often used on world scale
= 1,000 TJ, energy values often used on country scale

353

TJ
GJ
MJ
kJ
1 kWh
1 MWh
1 GWh
1 TWh
1 MJ
1 TOE
1 Barrel
a
h
ton
kton
M
G
T
P
E

354

= 1,000 GJ, energy value often used on plant scale


= 1,000 MJ, energy value often used on household scale
= 1,000 kJ
= 1,000 J
= 3.6 MJ
= 3.6 GJ
= 3.6 TJ
= 3.6 PJ
= 0.278 kWh
= 41.868 GJ
= 158.9873 litres
= annual
= hour
= 1,000 kg
= 1,000 ton
= Mega,1 x 106, 1 million
= Giga, 1 x 109, 1 billion
= Tera 1 x 1012
= Peta 1 x 1015
= Exa, 1 x 1018

Dankwoord
Na zeven jaar van onderzoek komt er met de afronding van dit proefschrift een einde aan een
bijzonder boeiende, leerzame en ook veeleisende periode. Energieconversie is een prachtig
vakgebied waarin de technologische ontwikkelingen snel gaan, gedreven door toekomstige
energieschaarste, brandstofprijsontwikkelingen en het klimaatprobleem. De snelheid van de
technologische ontwikkelingen, het feit dat we aan het begin staan van een vergaande verduurzaming
van het Nederlandse energie systeem en persoonlijke interesse in alles wat met energieconversie te
maken heeft, was reden voor dit promotieonderzoek.
Vele mensen zijn betrokken geweest bij mijn onderzoek en hebben op een of andere wijze
bijgedragen aan de realisatie van het proefschrift. Zonder volledig te kunnen zijn wil ik een aantal van
hen in het bijzonder bedanken.
Ad Verkooijen wil ik bedanken voor het gestelde vertrouwen en zijn begeleiding gedurende het
promotietraject. Het positieve en opbouwend commentaar tijdens elk overleg dat we hadden gaf mij
altijd weer veel inspiratie om met volle moed verder te werken aan de opbouw en invulling van het
proefschrift en de realisatie van het simulatie model.
Het onderzoek is gestart als duo onderzoek samen met Teus van Eck waarbij Teus het accent heeft
gelegd op de politiek-economische waardeketen van de energievoorziening. We zijn bijna 6 jaar zeer
intensief met elkaar opgetrokken en hebben veel discussies en overleg gevoerd over mijn en zijn
promotiewerk. Uiteindelijk hebben beide onderzoeken geresulteerd in twee boekjes die elkaar prima
aanvullen en veel relevante onderzoeksinformatie geven over de energievoorziening in al zijn
facetten. Van Teus zijn kennis en praktijk ervaring heb ik gedurende het onderzoek veel geleerd en
het heeft mijn blik op alles wat energie te maken heeft verruimd en daarvoor ben ik hem zeer
erkentelijk.
Bij de ontwikkeling van het simulatie model heb ik veel ondersteuning gehad van Jan Kromhout van
E.ON Benelux. Voor het op de juiste wijze invulling geven aan de benodigde optimalisatieroutine,
waarmee de inzet van elektriciteitscentrales wordt bepaald, was Jan door zijn jaren ervaring met
simulaties van onschatbare waarde voor mij. Ik heb het altijd erg gewaardeerd dat we in alle openheid
en op basis van wederzijds respect en vertrouwen met elkaar konden overleggen over mijn
onderzoek. Voor al zijn ondersteuning ben ik hem veel dank verschuldigd.
Dirk Jansen, ten tijde van mijn onderzoek directeur van Nuon Warmte, wil ik bedanken voor de ruimte
die ik heb gekregen om naast mijn reguliere werkzaamheden voor Nuon Warmte dit onderzoek te
mogen doen. Het gestelde vertrouwen in mij en het onderzoek betekent veel voor mij.
De ruimte en middelen die Nuon gegeven heeft om dit onderzoek mogelijk te maken is geweldig en
door mij zeer gewaardeerd. Het geeft temeer aan dat Nuon een bedrijf is waar de persoonlijke
ontwikkelingen van haar werknemers ondersteund worden, en dat veel mogelijkheden worden
geboden om jezelf te kunnen ontplooien.
Er zijn, zoals vermeld hierboven, veel mensen betrokken geweest bij dit onderzoek van wie ik er een
aantal nog expliciet wil noemen en bedanken.
Mijn collegas Bart Dijkman, Peter Buskermolen en Jan van den Bor van Energy Sourcing, Menno van
der Horst, Edward Nagelhout en Annieck Vennegoor van Warmte, voor de beantwoorden van mijn
vragen op het gebied van de energievoorziening, de bereidheid mijn gegevens, benodigd voor het
onderzoek, compleet te krijgen en voor de controle van teksten.
Rob Roumen voor het vertalen van de door mij aangeleverde stukken en het altijd stipt nakomen van
de afspraken.
Rolf Knneke voor zijn ondersteuning aangaande de opbouw van het proefschrift.
William Dhaeseleer voor het zeer nauwgezet bestuderen en van aanvullingen en commentaar
voorzien van het manuscript van het proefschrift. Dit is het uiteindelijke resultaat zeker ten goede
gekomen.

355

Familie, vrienden, kennissen en collegas wil ik bedanken voor de belangstelling die ze afgelopen
jaren hebben getoond voor het onderzoek en de voortgang daarvan.
Mijn ouders die me altijd hebben gesteund in al mijn keuzes. Mijn broer Steffan aan wie ik veel steun
heb gehad de afgelopen jaren.
Tot slot, van groot belang is de ondersteuning geweest van mijn vrouw Angelique en dochter Laurine.
Het zijn toch flink wat jaartjes geworden waarin veel is gebeurd en ik met dit onderzoek bezig ben
geweest. Door de juiste positieve invulling en jullie betrokkenheid hebben de gezinsactiviteiten er niet
onder hoeven te leiden. Bedankt voor al jullie steun de afgelopen jaren.

356

Curriculum Vitae
Hans Rdel was born in 1967 in Doetinchem the Netherlands. In 1991 he received the Bachelors
degree in Mechanical Engineering form the Technical University in Arnhem. In 1994 he received the
Masters degree in Mechanical Engineering, with a specialization in Thermal Engineering, from the
University of Twente (UT).
After receiving his Masters and a short period of military services Hans started to work for KEMA in
1995 as a scientific employee for the business unit Fossil Generation. In 1996 he joined Thomassen
International in Rheden at the subsidiary company Thomassen Power Systems. He started his career
there as process engineer of medium and large scaled Combined Heat and Power plants based on
General Electric gas turbines. After Thomassen Power Systems he started in 1998 to work for Nuon in
several functions and actual he works at the business unit Heat which covers all the District Heating
projects of Nuon. He is responsible for the purchase of heat, electricity and gas and the economic and
environmental optimisations of the existing and new to build heat- and cooling projects. From 2001 to
2008 he worked part time on a PhD study about Possible Future Scenarios for the Dutch Electricity
Supply System.

357

358

You might also like