dejong 2005 phdthesis

306
Nitrogen compounds in pressurised fluidised bed gasification of biomass and fossil fuels Wiebren de Jong

Upload: wiebrendejong

Post on 11-Apr-2015

1.068 views

Category:

Documents


8 download

DESCRIPTION

Nitrogen compounds in pressurised fluidised bed gasification of biomass and coal

TRANSCRIPT

Nitrogen compounds in pressurised fluidised bed gasification of biomass and fossil fuels

Wiebren de Jong

Nitrogen compounds in pressurised fluidised bed gasification of biomass and fossil fuels

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 maandag 7 februari 2005 om 15:30 uur

door Wiebren DE JONG scheikundig ingenieur geboren te Rotsterhaule (Haskerland)

Dit proefschrift is goedgekeurd door de promotoren: Prof. Dr. Ing. K.R.G. Hein Prof. dr. J.A. Moulijn Samenstelling promotiecommissie: Rector Magnificus, Prof. Dr. Ing. K.R.G. Hein, Prof. dr. J.A. Moulijn Prof. Dr. Ing. H. Spliethoff Prof. dr. ir. P.J. Jansens Prof. dr. ir. C. Daey Ouwens Prof. dr. M. Hupa Dr. ir. P.D.J. Hoppesteyn voorzitter Technische Universiteit Delft, promotor Technische Universiteit Delft, co-promotor Technische Universiteit Delft Technische Universiteit Delft Technische Universiteit Eindhoven bo Akademi University, Finland Corus, IJmuiden

Drs. J. Andries heeft als begeleider in belangrijke mate aan de totstandkoming van het proefschrift bijgedragen.

Published and distributed by: Optima Grafische Communicatie Pearl Buckplaats 37 3009 CC Rotterdam The Netherlands Telephone: +31 10 220 11 49 Telefax: +31 10 456 63 54 E-Mail: [email protected] ISBN 90-8559-028-0 Copyright 2005 by W. de Jong All rights reserved. No part of the material protected by this copyright notice may be reproduced or utilized in any form or by any means, electronic or mechanical, including photocopying, recording or by any information storage and retrieval system, without written permission from the publisher. Printed in the Netherlands

Alles heeft zijn uur en ieder ding onder de hemel zijn tijd. Prediker 3:1. There is a time for everything, and a season for every activity under heaven. Ecclesiastes 3:1.

Voor: Heit en mem en mijn echtgenote Klarine

Nitrogen compounds in pressurised fluidised bed gasification of biomass and fossil fuels

SummaryThis PhD thesis assesses the experimental and theoretical work which was performed to study the behaviour of nitrogen compounds during airblown pressurised stationary fluidised bed gasification of biomass and brown coal, followed by high temperature ceramic gas filtration. Fossil fuels have dominated the energy supply in modern societies and will continue to do so in the 21st century. The resources, however, are depleting, especially of oil and natural gas. Therefore, other energy sources are to be exploited further within this century. Biomass is one of the almost CO2 neutral, renewable contributors to the future energy production. Nowadays many modern, high efficiency (combined) power and heat producing systems using biomass are or become commercially available. One promising route to efficient power and heat supply is the Integrated Gasification Combined Cycle. This cycle is particularly of interest for medium to larger scale installations. Pressurised operation of the gasifier offers the advantage of smaller process equipment, including that for the necessary downstream gas cleaning. Also, the work needed to compress the gas for gasturbine use will be much smaller or not needed at all. High temperature gas filtration offers the benefit of increased overall efficiencies of the power and heat producing cycle. This integrated gasification technology, however, is still in a stage of development and demonstration. When instead of absorption gas cleaning, high temperature, dry gas filtration is applied, nitrogen compounds, like ammonia (NH3) and hydrogen cyanide (HCN), are not dissolved in the absorption liquid. As a result NOx emissions in gas turbine combustors are produced. NOx is known for its negative effects on the health of humans and animals and acidification of soil and water. Therefore, increasingly stringent emission restrictions are imposed on this component. Both coal and biomass contain nitrogen in their chemical structure and in gasification processes this so-called fuel bound nitrogen is converted to a large extent into NOx precursors. Although woody biomass contains low amounts of nitrogen (only a few tens of mass percentage on dry fuel basis), there is still a high emission potential based on the fuel's energy content, because of the low calorific value of the fuel, as compared to coal. Thus biomass causes significant NOx emissions when no further measures are taken. An introduction regarding the use of biomass in energy production, the potential of NOx emissions of a range of young and old fuels and open research questions is given in chapter 1. A literature overview is presented in chapter 2. A state-of-the-art review of the fluidised bed gasification activities is given. Also, an overview of the modelling of fluidised bed gasification on large and small scale, with emphasis on the emission of NOx precursors is presented. The influence of fuels, additives and process parameters on the release of these compounds is addressed. Both the primary and later stages of the conversion of solid fuels are considered. In chapter 3 the experimental facilities used to study the thermal conversion behaviour of fuel bound nitrogen are presented. These facilities can be divided into two categories: 1) Pilot scale test rigs; a 1.5 MWthermal Pressurised Fluidised Bed Gasifier situated at the section Energy Technology of Delft University of Technology and a similar 50 kWthermal test rig available at the Institute of Process Engineering and Power Plant Technology of Stuttgart University (Germany).

vii

2) Characterisation equipment to study solid fuel reactivity, especially in the early stages of conversion in gasification processes, namely pyrolysis. Here, two techniques are used representing slow and rapid heating conditions, respectively: TG-FTIR (thermogravimetric analysis coupled to FTIR), situated at AFR inc. (Hartford, CT, USA) and heated grid reactor equipped with in-situ IR diode laser diagnostics, available at the Technical Physics department of Eindhoven University of Technology. In chapter 4 the experimental results obtained by using the facilities described above are presented and discussed. In the pressurised fluidised bed gasification tests performed at Delft university, miscanthus and wood pellets have been used. Brown coal has been chosen because it is an older, but still comparatively high-volatile, fossil fuel. No significant radial gradients in the concentration of the main and minor gaseous product constituents were observed. The concentrations of the main gasification product gas components were comparable to the limited open literature data, available from other pressurised fluidised bed test rigs at comparable air stoichiometry values. Axial gradients in the gas concentrations during the pressurised fluidised bed gasification tests could be clearly observed for acetylene, which is related to reactions involving tar and soot precursor formation and destruction. Under the pressurised fluidised bed gasification conditions studied, the main fuel bound nitrogen component produced is NH3, whereas HCN is formed to a minor extent (only a few percent of the fuel bound nitrogen). This was comparable with results from other bottom-fed FB gasifiers. On the other hand, comparatively low values have been found for a top-fed pressurised FB. HNCO and NO were never detected by means of even a high resolution FTIR spectrophotometer under the pressurised gasification test conditions studied. Tests with Ca-containing dolomite and a Ca-less additive (MinPhyl, or Pyrophyllite) under otherwise comparable process conditions showed that an increased Ca inventory in the gasifier increases the NH3/HCN ratio significantly. To obtain basic model input data, flash pyrolysis experiments with miscanthus were conducted using the heated grid reactor set up. This research was focused on measuring the yield of CO, CO2 and NH3 at a heating rate of 280-320 K/s and final temperatures in the range of 1050-1400K. Unfortunately, NH3 could not be detected, due to condensation or the limited frequency range that could be achieved with the tuneable laser. CO and CO2 yields have been measured in-situ and were compared with the FG-DVC biomass pyrolysis model. For all fuels used, kinetic parameters for this pyrolysis model have been determined by application of the Tmax method, using a TG-FTIR system with heating rates of 10, 30 and 100 K/min. However, this model does not predict the pyrolysis product yield satisfactorily at high heating rates, based on the kinetics determined by low heating rates. It gave a reasonable quantitative yield prediction for CO but a substantial under-prediction for CO2. The competition between the evolution of primary products like primary tar fragments and carboxylic acids on one side and light gases like CO, CO2 and H2O on the other side is probably the reason. This competition is expected to be heating rate dependent. Apparently the primary pyrolysis products like tars and carboxylic acids, which contain precursor groups for the formation of CO and CO2, are quickly removed from the reaction zone and quenched immediately. Thus, no time is available for further primary tar and carboxylic acid decomposition into CO2 and CO (to a minor extent), which results in low yields. According to this hypothesis the yields of primary tar and carboxylic acids must be significant. This is confirmed by the observation that under pyrolysis conditions the tar yield increases at increasing heating rates. In chapter 5 the modelling of pressurised fluidised bed gasification is described and the influence of process parameters is assessed. The model is a steady state plug flow-in-series model with detailed reaction kinetics. Heterogeneous char oxidation and gasification (by H2O, CO2 and H2), heterogeneously catalysed HCN hydrolysis and homogeneous reactions (including nitrogen molecular and radical species and a simplified tar cracking reaction) are taken into account.

viii

The possibility to increase the conversion of NH3 into N2 by varying the process conditions or by adding specific compounds to the gasifier has been studied theoretically. However, NH3 is found to be a very stable compound, which is hardly converted in pressurised fluidised bed gasifiers. Decomposition into N2 is slightly increased by increasing temperature, but this option is limited due to the risk of bed sintering for alkali containing biomass fuels. The NH3 conversion is only slightly dependent on reactor pressure. A minimum conversion was obtained around 2.5 bar. At higher pressures (10 bar) the NH3 conversion slightly increased. Gas residence time in the reactor did not affect the fuel-N conversion. The destruction of NH3 is only taking place in the presence of O, H and OH radicals, which are consumed very fast in the initial part of the bed. Larger NH3 conversions can be reached by injecting NO or NO2 into the bed. However, HCN is formed and NO conversion is limited, which leads to undesired emissions. Addition of O2 favours NH3 conversion. However, the main nitrogen species formed was NO rather than N2. In the most favourable case, at near-stoichiometric conditions, NH3 is converted for 50% into NO and 50% into N2. Moreover, the addition of secondary air decreases the already low LCV gas heating value. The presence of a high concentration of CH4 in the bed part of the gasifier reduces the NH3 conversion, probably due to the competition for radicals between CH4, its intermediates (mainly CH3 radicals) and NH3. H2O2 and H2O (steam) addition into the bed did not affect at all NH3 conversion. A comparison between the model based simulations and the experiments on the 1.5 MWth Delft University pressurised fluidised bed gasification scale and the 50 kWth Stuttgart University rig is presented in chapter 6. The agreement for N-species prediction and measurements is quite good for the fuels and more in particular for the main fuel bound nitrogen component in the product gas: NH3. For HCN the concentrations are often underpredicted, probably due to the heterogeneous hydrolysis reaction, taking place at the char surface, which can have slower kinetics than assumed in literature. The model predicts the formation of super-ppmv HNCO and NO concentrations, but they have never been detected by FTIR analysis in our work. Probably, catalytic hydrolysis converts HNCO into NH3 and CO and this reaction has not been considered in the model. For NO, catalytic reduction by ash constituents probably plays a role at the gasification temperatures. It is also possible that neglection of S and Cl chemistry causes deviations between model and experimental results for the minor species. This is less the case for wood. Also, simplification of tar- and char-nitrogen reactions can be significant: the model assumes that the nitrogen which is not available as gas species will be released initially in the form of HCN. The agreement between model and experimental results for the main product gases is reasonably good. The differences between the calculated and measured values can be attributed to pyrolysis yields of H2. These were obtained from a correlation of literature data and as such used as input in the biomass pyrolysis sub-model. Also, the use of simplified tar cracking kinetics, in terms of possible reactions, product yields and rates probably plays a major role in the deviations observed for CO and CO2 concentrations. The deviations are the highest for wood, and less for miscanthus and brown coal. This is in-line with the hypothesis that the simplified tar cracking kinetics plays a major role, as of the tested fuels wood pyrolysis shows the highest initial tar yields. Differences in the heterogeneous combustion and gasification reactions are not expected to have a major impact, as the carbon conversions are quite well predicted. Finally, chapter 7 gives an overview of the conclusions and the recommendations specified for pressurised fluidised bed experimental pilot scale research, fuel characterisation and modelling.

Wiebren de Jong

ix

Stikstofcomponenten in drukwervelbedvergassing van biomassa en fossiele brandstoffen

SamenvattingDit proefschrift behandelt het experimentele en theoretische werk dat is uitgevoerd in het kader van een studie naar het lot van de stikstof componenten tijdens drukwervelbedvergassing in een stationair wervelbed in combinatie met hoge temperatuur filtratie, gebruikmakend van keramische filters. Fossiele brandstoffen hebben de energievoorziening van de moderne maatschappij bepaald en zullen dit blijven doen in de 21e eeuw. De voorraden raken echter uitgeput, in het bijzonder die van olie en aardgas. Daarom moeten andere energiebronnen in deze eeuw worden aangeboord. Biomassa is een van de bijna CO2 neutrale, hernieuwbare brandstoffen voor de toekomstige energievoorziening. Tegenwoordig zijn er al commercile systemen op de markt voor (gecombineerde) warmte en kracht productie met een hoog rendement op basis van biomassa. Een veelbelovende optie voor efficinte elektriciteit- en warmtevoorziening is het gecombineerde vergasser STEG systeem. Dit systeem is in het bijzonder van belang voor middelgrote tot grote installaties. Het onder druk bedrijven van de vergasser biedt het voordeel van kleinere proces apparatuur, inclusief de downstream gasreiniging. Compressie van het geproduceerde gas, hetgeen nodig is voor het bedrijven van moderne gasturbines, zal in dit geval niet of in mindere mate nodig zijn. Hoge temperatuur gasreiniging door middel van bijvoorbeeld keramische filters biedt het voordeel van hogere efficinties van het warmte en elektriciteit producerende systeem. Deze gentegreerde vergassingstechnologie, echter, bevindt zich nog steeds in het stadium van ontwikkeling en demonstratie. Bij toepassing van hoge-temperatuur en dus droge gas filtratie in plaats van natte absorptie technieken zullen stikstofcomponenten, zoals ammoniak (NH3) en waterstofcyanide (HCN), niet oplossen in de absorptievloeistof en zullen NOx vormen in gasturbine verbrandingskamers. NOx heeft negatieve effecten op de gezondheid van mens en dier en veroorzaakt verzuring van de grond en het oppervlaktewater. Daarom worden wereldwijd steeds stringentere emissie eisen opgelegd voor deze emissiecomponent. Zowel kolen als biomassa bevatten chemisch gebonden stikstof en het is deze brandstofgebonden stikstof die in vergassingsprocessen voor een groot deel wordt omgezet in NOx precursors. Hoewel houtachtige biomassa lage gehaltes aan stikstof vertoont (slechts enkele tiendes massaprocenten op droge basis), is er in vergelijking met kolen toch een hoog emissiepotentieel als de lage stookwaarde van de brandstof in ogenschouw wordt genomen. Daarom draagt biomassa significant bij tot NOx emissies wanneer er geen proces gerelateerde maatregelen worden genomen. In hoofdstuk 1 van dit proefschrift wordt een inleiding gegeven in de toepassing van biomassa in de energievoorziening, potentile NOx emissies van een reeks jonge tot oude brandstoffen, alsmede open onderzoeksvragen op dit gebied. In hoofdstuk 2 wordt een literatuuroverzicht geboden. Hierin wordt een state-of-the-art overzicht gegeven van de wervelbedvergassingsactiviteiten op zowel kleine als grote schaal en wordt een overzicht gepresenteerd van de modellering van wervelbedvergassing. De nadruk wordt gelegd op de emissie van NOx precursors, zowel in de primaire als ook de latere stadia van de conversie van vaste brandstoffen. De invloed van verschillende brandstoffen, additieven en proces gerelateerde parameters op het vrijkomen van deze componenten wordt behandeld. Hoofdstuk 3 geeft een overzicht van de toegepaste experimentele technieken de gebruikt zijn voor de studie naar het thermische conversiegedrag van brandstofgebonden stikstof. Deze kunnen worden onderverdeeld in twee categorien:

xi

1) Proefopstellingen op pilot schaal: de 1.5 MWthermisch drukwervelbedvergassingsopstelling in het laboratorium van de sectie Energy Technology van de Technische Universiteit Delft, en de 50 kWthermisch opstelling die beschikbaar is bij het Institut fr Verfahrenstechnik und Dampfkesselwesen (IVD) van de Universiteit Stuttgart (Duitsland). 2) Fundamentele karakteriseringsapparatuur voor bestudering van de vaste brandstofreaktiviteit, in het bijzonder in de vroege stadia van conversie in het vergassingsproces, namelijk de pyrolyse. Hier worden twee technieken besproken die langzame en snelle verhittingscondities vertegenwoordigen: respectievelijk TG-FTIR (thermogravimetrische analysie gekoppeld met FTIR), bij AFR Inc. (Hartford, CT, USA) en de heated grid reactor voorzien van in-situ IR diode laser diagnostiek, bij de afdeling Technische Natuurkunde van de Technische Universiteit Eindhoven. In hoofdstuk 4 worden de experimentele resultaten gepresenteerd en besproken. In de Delftse drukwervelbed experimenten zijn miscanthus- en houtpellets als biobrandstoffen gebruikt. Bruinkool is geselecteerd omdat het een oudere, maar toch nog hoog-vluchtige fossiele brandstof is. Er is geen significante gradint in het radiaal concentratieprofiel van hoofd- en sporecomponenten waargenomen in de vergasser. De concentraties van de hoofdcomponenten van het productgas waren vergelijkbaar met de beperkte gemeten open literatuurdata van andere drukwervelbed opstellingen bij vergelijkbare stoichiometrie. In de axiale gas concentratie profielen tijdens de Delftse drukwervelbed experimenten werden duidelijke gradinten waargenomen voor acetyleen, een component welke gerelateerd is aan reakties die teer en roet-precursor vorming en afbraak betreffen. Bij de toegepaste drukwervelbed vergassingscondities is de belangrijkste brandstofgebonden stikstof component NH3, terwijl HCN in mindere mate wordt gevormd (enkele procenten van de vaste brandstofgebonden stikstof). De conversie naar NH3 en HCN was vergelijkbaar met andere FB vergassers met bodemvoeding, in contrast met de relatief lage waardes die werden waargenomen voor een drukwervelbed met topvoeding. HNCO and NO zijn nooit gedetecteerd, zelfs niet met een hogeresolutie FTIR spectrofotometer, onder de bestudeerde drukvergassingscondities. Experimenten uitgevoerd met Ca-houdend dolomiet en een Ca-loos additief (MinPhyl, of Pyrophylliet) onder vergelijkbare procescondities toonden aan dat een verhoogd Ca aanbod in de vergasser leidt tot een significante toename in de NH3/HCN verhouding. Ter verkrijging van de basis model inputgegevens, zijn heated grid flash pyrolyse experimenten uitgevoerd met miscanthus. Dit deelonderzoek was gericht op het bepalen van opbrengst van CO, CO2 and NH3 als functie van de temperatuur in de range 1050-1400K en opwarmsnelheden van 280320K/s. NH3 kon helaas niet worden gedetecteerd, hetgeen te wijten was aan ofwel condensatie of het beperkte frequentie gebied dat kon worden ingesteld met de tuneable laser. De CO en CO2 opbrengsten zijn in-situ bepaald en vergeleken met de FG-DVC biomassa pyrolyse model uitkomsten. Voor alle toegepaste brandstoffen zijn de kinetische parameters bepaald voor dit pyrolyse model door toepassing van de Tmax methode, uitgaande van TG-FTIR metingen bij verhittingssnelheden van 10, 30 en 100 K/min. De extrapolatie die dit model gebruikt om de pyrolyse productopbrengst bij hoge opwarmsnelheden te voorspellen gebaseerd op de kinetiek parameters die verkregen zijn bij lage opwarmsnelheden, levert een wisselend beeld op. Het resulteert in een redelijk correcte kwantitatieve opbrengstvoorspelling voor CO en een duidelijk te lage voorspelling voor CO2. De competitie tussen het vrijkomen van primaire producten als primaire teerfragmenten en carboxylzuren enerzijds en lichte gassen als CO, CO2 en H2O anderzijds is waarschijnlijk de oorzaak. Deze competitie is waarschijnlijk afhankelijk van de opwarmsnelheid. De primaire pyrolyse producten als teren en carboxylzuren, die precursor groepen bevatten voor CO en CO2 vorming, worden blijkbaar snel uit de reaktiezone verwijderd en direct afgekoeld. Er is dan dus geen tijd beschikbaar voor verdere ontleding in CO2 en (in mindere mate) CO, wat resulteert in lage productopbrengsten.

xii

Volgens deze hypothese moeten de opbrengsten van primaire teren en carboxylzuren significant zijn. Dit wordt bevestigd door de waarneming dat voor pyrolyse bij toenemende opwarmsnelheden de teer opbrengst toeneemt. In hoofdstuk 5 wordt de modellering van drukwervelbedvergassing behandeld. Hier wordt het model beschreven en de invloed van proces parameters bestudeerd. Het model is een stationair propstroomreactor-in-serie model met gedetaileerde reactiekinetiek. Heterogene kool-residue oxidatie en vergassing (door H2O, CO2 en H2), heterogeen gekatalyseerde HCN hydrolyse en homogene reacties (met inbegrip van stikstofhoudende moleculen en radicalen en een vereenvoudigde teer kraakreactie) worden in het model meegenomen. De mogelijkheid om NH3 conversie in N2 te vergroten door procescondities te variren of door toevoeging van specifieke componenten aan de vergasser is theoretisch bestudeerd. Het blijkt dat NH3 een erg stabiele component is, welke nauwelijks kan worden omgezet in drukwervelbedvergassers. Decompositie in N2 neemt iets toe door verhoging van de temperatuur, hoewel deze optie beperkt is door het risico van bed sintering voor alkali-houdende biobrandstoffen. De NH3 conversie is slechts in beperkte mate afhankelijk van de druk. Een minimum in de conversie is gevonden bij 2.5 bar. Bij hogere drukken (10 bar) neemt de NH3 conversie licht toe. De gas verblijftijd in de reaktor heeft praktisch geen invloed op de brandstof-N conversie. Afbraak van NH3 vindt alleen plaats in de aanwezigheid van O, H en OH radicalen, die erg snel reageren in de initile bedzone. Hogere NH3 conversies kunnen worden gerealiseerd door injectie van NO of NO2 in het bed. Anderzijds wordt dan HCN gevormd en wordt ongereageerd NO voorspeld, die beide ongewenste emissiecomponenten zijn. Additie van O2 bevordert de NH3 conversie. De belangrijkste stikstofcomponent wordt dan echter NO en in het meest ideale geval, bij praktisch stoichiometrische condities, wordt de brandstofgebonden N voor 50% in NO omgezet en voor 50% in N2. Toevoegen van secundaire lucht verlaagt de toch al lage LCV gas stookwaarde. Aanwezigheid van een hoge concentratie CH4 in de bedzone van de vergasser reduceert de NH3 conversie, waarschijnlijk door de competitie voor radikalen tussen CH4, haar intermediaire omzettingsproducten (voornamelijk CH3 radikalen) en NH3. H2O2 and H2O (stoom) additie in het bed heeft geen invloed op de NH3 conversie. Een vergelijking tussen de met behulp van het in hoofdstuk 5 beschreven model uitgevoerde simulaties en de experimenten uitgevoerd op de 1.5 MWth drukwervelbed schaal en de 50 kWth schaal wordt in hoofdstuk 6 gepresenteerd. De overeenkomst tussen berekende en gemeten waarden is voor de stikstofcomponenten tamelijk goed voor de brandstoffen en in het bijzonder voor de belangrijkste brandstofgebonden stikstofcomponent: NH3. Voor HCN wordt de concentratie vaak te laag voorspeld, mogelijk door de heterogeen gekatalyseerd hydrolyse reaktie, welke plaatsvindt op het kool-residue oppervlak en die een langzamere reaktiekinetiek zou kunnen hebben dan aangenomen is in de literatuur. Het model voorspelt ruim 10 ppmv HNCO en NO, maar deze stoffen zijn nooit door middel van FTIR analyse gedetecteerd. Waarschijnlijk wordt HNCO katalytisch in NH3 en CO omgezet en deze reaktie is niet in het model meegenomen. Voor NO speelt katalytische reductie door as-elementen waarschijnlijk een rol bij de heersende vergassingstemperaturen. Het is ook mogelijk dat verwaarlozing van de S en Cl chemie voor de trace componenten de afwijkingen tussen model en experimenteel resultaten veroorzaakt. Dit is in mindere mate het geval voor hout. De vereenvoudiging van de teer- en koolresiduegebonden stikstof kan ook van betekenis zijn: in het model wordt aangenomen dat de stikstof die niet vrijkomt als gas component bij de pyrolyse, initieel als HCN vrijkomt. Voor de hoofdcomponenten van het productgas is de overeenkomst tussen de model- en experimentele resultaten tamelijk goed. De verschillen tussen de berekeningen en de metingen kunnen worden toegeschreven aan de op basis van de literatuur aangenomen pyrolyse opbrengst van waterstof in het biomassa pyrolyse sub-model. De onzekerheid aangaande de teerontledingskinetiek, in termen van mogelijke reacties, product opbrengsten en snelheid, speelt waarschijnlijk ook een prominente rol in de waargenomen verschillen voor CO en CO2.

xiii

De afwijkingen zijn het hoogst voor hout en minder voor miscanthus en bruinkool. Dit komt overeen met de hypothese dat onnauwkeurigheden in de teerontledingskinetiek hier een belangrijke rol in spelen, omdat hout de hoogste initile teeropbrengst vertoont in de snelle pyrolyse stap. Verschillen in de heterogene verbrandings- en vergassingsreactiekinetiek hebben naar verwachting een veel minder grote invloed op de voorspelde concentraties, aangezien de koolstofconversie in het algemeen tamelijk goed wordt voorspeld. Tenslotte wordt in hoofdstuk 7 een overzicht gegeven van de getrokken conclusies en aanbevelingen gedaan voor verder onderzoek, gespecificeerd naar experimentele drukwervelbed pilot schaal onderzoek, brandstofkarakaterisering en modellering.

Wiebren de Jong

xiv

Table of ContentsSummary Samenvatting Notation Chapter 1: Introduction1.1 1.2 1.3 Towards a renewable energy based world Biomass as part of renewable power generation Technology for biomass utilisation for heat and power generation 1.3.1 Small scale versus large scale processes 1.3.2 Technological options for large scale biomass based heat and power generation 1.3.2.1 Combustion 1.3.2.2 Gasification Power production from biomass gasification, open research questions The fate of fuel bound nitrogen Outline of this thesis 1 3 3 4 5 5 6 8 9 11 vii xi xix

1.4 1.5 1.6

Chapter 2: Fluidised bed solid fuel gasification processes, overview and analysis of experimental research and modelling2.1 2.2 2.3 2.4 The fluidised bed reactor applied for solid fuel gasification Industrial fluidised bed gasification systems An overview of recent research, development and small scale demonstration activities Experimental findings regarding the fate of fuel nitrogen during fluidised bed gasification 2.4.1 Influence of fuel type 2.4.2 Influence of air stoichiometry 2.4.3 Influence of temperature 2.4.4 Influence of pressure 2.4.5 Influence of additives 2.4.6 Influence of particle diameter 2.4.7 Influence of steam 2.4.8 Influence of feed location Fluidised bed gasifier modelling 2.5.1 General overview 2.5.2 Drying and flash pyrolysis, initial steps in the process 2.5.2.1 Experimental techniques and findings 2.5.2.1.1 Main components and hydrocarbons 2.5.2.1.2 Nitrogen components 2.5.2.2 Modelling approaches 2.5.3 Heterogeneous char-gas reactions 2.5.3.1 Main carbon based reactions 2.5.3.2 Heterogeneous and heterogeneously catalysed homogeneous nitrogen reactions 2.5.4 Homogeneous gas phase reaction mechanisms, including nitrogen chemistry Potential primary measures for fuel_NOx emission reduction Conclusions and research requirements 13 13 17 20 20 23 24 25 26 26 27 27 28 28 32 32 33 53 60 62 62 65 66 68 70

2.5

2.6 2.7

xv

Chapter 3: Experimental set-ups and measurement techniques3.1 3.2 Introduction The Delft Pressurised Fluidised Bed Gasification (PFBG) test rig 3.2.1 Description of the rig 3.2.2 Analysis and sampling techniques 3.2.2.1 FTIR spectrophotometer 3.2.2.2 Gas Chromatography 3.2.2.3 On-line Non Dispersive Infrared/UV, colorimetric and paramagnetism based analysers 3.2.2.4 Sampling probes and analysis of tar compounds The 50 kW(thermal) IVD Pressurised Fluidised Bed Gasifier (DWSA) 3.3.1 Description of dimensions and operation 3.3.2 Analysis techniques applied The TG-FTIR set-up at Advanced Fuel Research Company Inc. (USA) The heated grid reactor at Eindhoven University of Technology 71 71 71 75 75 81 83 86 89 89 90 91 92

3.3

3.4 3.5

Chapter 4: Experimental results4.1 Choice of fuels, bed materials and additives 4.1.1 Fuel choice 4.1.2 Fuel composition and related chemical properties 4.1.3 Physical property characterisation of fuels and bed materials Experimental results of PFBG gasification tests 4.2.1 Experimental data representation and definitions of relevant parameters 4.2.2 Background information on the measurement campaigns 4.2.3 Miscanthus gasification 4.2.4 Wood gasification 4.2.5 Brown coal gasification Experimental results of DWSA gasification tests 4.3.1 Overview of the DWSA measurement programme 4.3.2 Wood gasification 4.3.3 Brown coal gasification Experimental results of TG-FTIR pyrolysis tests 4.4.1 Overview of the TG-FTIR experimental programme 4.4.2 Kinetic analysis approach 4.4.3 TG-FTIR analysis results and derived kinetic parameters for miscanthus 4.4.4 TG-FTIR analysis results and derived kinetic parameters for wood (Labee A quality energy pellets) 4.4.5 TG-FTIR analysis results and derived kinetic parameters for brown coal 4.4.6 TG-FTIR analysis discussion Experimental results of heated grid pyrolysis tests 4.5.1 The heated grid experimental programme 4.5.2 Miscanthus pyrolysis results 4.5.3 Discussion of the results Conclusions and recommendations 4.6.1 Conclusions and recommendations related to PFB gasification 4.6.2 Conclusions and recommendations for fuel characterisation 99 99 100 101 104 104 105 106 117 124 129 129 130 131 133 133 133 135 140 144 147 149 149 149 153 155 155 156

4.2

4.3

4.4

4.5

4.6

xvi

Chapter 5: Modelling bubbling fluidised bed gasification, focussed on nitrogen compounds5.1 5.2 5.3 Modelling approach Description of the model 5.2.1 Idealised reactor approach Simulation results using the idealised reactor modelling approach Conclusions 159 159 159 169 182

5.4

Chapter 6: Comparison of modelling results and experiments6.1 6.2 6.3 6.4 Choices made for the comparison between model and experiments Gasification experiments compared with model results 6.2.1 TUD PFBG experimental and simulation results 6.2.2 IVD DWSA experimental and simulation results Discussion of the results Conclusions and recommendations 183 183 183 192 196 197

Chapter 7: Conclusions and recommendations for further research7.1 7.2 Conclusions Recommendations 199 201 203

Bibliography Appendices1 Details of analytic measurements1.1 Spectra, spectral windows and calibration curves used for quantitative species analysis with FT-IR 1.2 Calibration curves used for quantitative species analysis with gas chromatography

227 227 241 249 249 250 251 252 255 259 275 277 281 283

2

Relevant chemical & physical properties of the gasification product gas components2.1 Gas phase viscosity 2.2 Diffusion coefficients of gas phase components 2.3 Gas Phase Thermal Conductivity 2.4 Thermodynamic data

3 4 5 6

Detailed homogeneous reaction scheme Kilpinen 97 Results of TG-FTIR measurements and comparison of FG-DVC model results with experiments at different heating rates Calculation of axial Pclet numbers for PFBG and DWSA tests simulated List of publications

Dankwoord Curriculum Vitae

xvii

NotationAbbreviationsABFB ACFB AFR ar or a.r. BC BHF BTX CFD daf db FBN FG-DVC FID FTIR GC Gtoe HCV or HHV HGR HRSG IGCC LCV or LHV mf MSW Mtoe NDIR NDUV ODE PFB PBFB PCFB PFBG RDF RPS SCADA SCO SCR SPA SPE SRC SS TCD TGA TUD WP Atmospheric Bubbling Fluidised Bed Atmospheric Circulating Fluidised Bed Advanced Fuel Research inc. As received Brown Coal Baghouse Filter Benzene, Toluene, Xylene Computational Fluid Dynamics Dry and ash-free basis Dry basis Fuel Bound Nitrogen Functional Group-Depolymerisation Vaporisation Condensation Flame Ionisation Detector Fourier Transform InfraRed Gas Chromatograph Gigaton oil equivalent Higher Calorific Value Heated Grid Reactor Heat Recovery Steam Generator Integrated Gasification Combined Cycle Low Calorific Value Moisture free Municipal Solid Waste Megaton oil equivalents Non Dispersive InfraRed Non Dispersive UltraViolet Ordinary Differential Equation Pressurised Fluidised Bed Pressurised Bubbling Fluidised Bed Pressurised Circulating Fluidised Bed Pressurised Fluidised Bed Gasifier Refuse-derived Fuel Rotating Particle Separator Supervision Control and Data Acquisition Selective Catalytic Oxidation Selective Catalytic Reduction Solid Phase Adsorption technique Solid Phase Extraction Short Rotation Coppice Stainless Steel Thermal Conductivity Detector Thermogravimetric Analysis (or Analyser) Delft University of Technology Wood Pellets

Latin SymbolsA A0 Area of gasifier reactor part Area of nozzle holes [m2] [m2]

xix

A() ai() Bo b b ci cp CC d Dax DT Di,m E f g Ga H H I ki ki0 mi M MWi OF OM P Pe R Re t T u v x X Yi

Absorbance at wave number Absorption coefficient at wave number of species i' Bodenstein number Path length through the sample Temperature coefficient in the Arrhenius equation Concentration of species i' in the sample Specific heat Carbon Conversion Diameter Axial dispersion coefficient Diameter of gasifier reactor part Diffusion coefficient of component i in mixture Activation energy Radiation frequency Acceleration of gravity (= 9.81) Galilei number Enthalpy Height of bed Intensity of radiation Rate constant of component i Frequency factor (or pre-exponential factor) Mass fraction of i in fuel Heating rate Mole mass of component i Distance between beamsplitter and fixed mirror Distance between beamsplitter and movable mirror Pressure Pclet number Universal gas constant (=8314.3) Reynolds number Time Temperature Velocity Velocity Fraction of reacted material Mass fraction of char remaining Mass fraction of component or functional group i

[-] [m-1] [-] [m] [-] [-] [J.kmol-1.K] [%] [m] [m2.s-1] [m] [m2.s-1] [J.kmol-1] [s-1] [m.s-2] [-] [J.kmol-1] [m] [*] [s-1] [s-1] [-] [K.s-1] [kg.kmol-1] [m] [m] [Pa] [-] [J.kmol-1.K-1] [-] [s] [K] [m.s-1] [m.s-1] [-] [-] [-]

Greek Symbols m Retardation, or difference in optical path length Volume fraction Dynamic viscosity Efficiency Wavelength Wave number Kinematic viscosity Density Parameter in the empirical relation for Gaussian distribution of the activation energies Sphericity factor bed material ass flow Reaction rate of component j [m] [-] [Pa.s] [%] [m] [cm-1] [m2.s-1] [kg.m-3] [] [J.kmol-1] [-] [kg.s-1] [kmol.m-3.s-1]

j

.

xx

IndicesA ax b bed C diff e f g m max mf mix p r r s SMD Stoich th x Absolute Axial direction Bubble Bed (section) Carbon (char) Diffusion Electrical Forward reaction Gas (phase) Mass Maximum At minimum fluidisation condition Mixture (gas) Particle Residence Backward reaction Solid (phase) Sauter Mean Diameter Stoichiometric Thermal In axial direction

xxi

Chapter 1Introduction1.1 Towards a renewable energy based world Up till now, fossil fuel utilisation has made a large contribution to the energy systems of our modern age. However, most of the fossil fuel reserves for power and heat supply, transportation and chemicals manufacturing decrease and need to be substituted by alternatives in order to maintain the present way of life. The 20th century was based on these conventional energy carriers, but the 21st century will have to adjust to the decline of the oil and natural gas based economy and the increasing public awareness of the negative effects of environmental pollution from fossil fuel utilisation. Figures 1.1a through 1.1d give an overview of recent data concerning fossil fuel and uranium reserves, as well as the expected cumulative demand over the period 1990-2050.57 Gtoe51 Gtoe

Uranium

Oil 343 Gtoe

Uranium

Oil 411 Gtoe

Coal 606 Gtoe

Coal 710 Gtoe

Natural Gas 333 Gtoe

Natural Gas 237 Gtoe

Figure 1.1a Global primary Energy Reserves, Commercially and technically exploitable, Global Energy Perspective, [IIASA/WEC, 1998].40 Gtoe Uranium

Figure 1.1b Global primary Energy Reserves Commercially and technically exploitable, World Economic Outlook, [IEA, 1998].

Oil 261 GtoeCoal 2.13 Gtoe

Uranium

0.65 Gtoe

Coal 273 Gtoe

Oil 3.46 Gtoe

Natural Gas 211 Gtoe

Natural Gas 2.06 Gtoe

Figure 1.1c Cumulative primary energy demand (1990-2050) Energy Reserves, [IIASA/WEC, 1998].

Figure 1.1d Primary energy utilisation in 1999, [BP Amoco, 2001].

Figures 1.1a-c show that reserves of oil and natural gas are (just) sufficient for the indicated period. Based on previous experience, however, it is expected that more reserves of natural gas will be found in e.g. hydrate fields (clathrates) in deep-sea areas, but the quantities still have to be investigated.

1

The exploitation of these fields, though, will be comparatively expensive. Also for oil there are probably unknown reserves, which are by now not yet commercially or technically exploitable. Coal as conventional fuel is expected to be still available for several centuries because of its huge reserves and its broad global reserve distribution. The figures further indicate that the reserves of Uranium, the fuel for power generation based on nuclear fission, are of the same order of magnitude as the expected demand in the indicated period. This situation is not positive for the long term use of this fuel, unless technology is shifted to a more efficient use of fuel sources, like e.g. in fast breeder systems [Gardiner, 1990]. However, radioactive waste disposal and removal problems are still preventing the widespread using of nuclear technology. Also, the danger of proliferation of nuclear energy knowledge and implementation on a global scale might lead to increased nuclear arms proliferation. Furthermore, the world is faced with a growing human population and an increasing level of industrialisation, giving rise to an increasing demand for personal comfort and energy demand per capita. Long term scenarios, like the one mentioned in figure 1.1 c, differ in the way the energy use per capita increases, but in all scenarios 50-60% of the worlds energy supply will rely on fossil sources as primary energy supply. The use of fossil fuels as primary energy source gives rise to an increasing emission of environmentally hazardous species, which is becoming a growing problem. The rising worldwide awareness of environmental constraints leads to legislative actions in most industrialised countries in order to restrict local and regional emissions of acid rain or smog precursor gases (e.g. SOx and NOx) and dust. Furthermore, several industrialised countries are heading for a reduction of the human contribution to the greenhouse effect, by reducing emission of these so-called greenhouse gases, like in particular CO2, CH4 and N2O. Of these gases, CO2 is the most important contributor to the absorption of infrared radiation emitted from the earths surface. The increase of the CO2 concentration in the atmosphere is supposed to contribute significantly to the enhanced greenhouse effect (from 55% [Wjtowicz et al., 1993] to 63.5% [Sloss, 2002]). CH4 also absorbs infrared radiation and contributes for approximately 15% -20.5% to the enhanced greenhouse effect. Finally, N2O has a contribution of about 6.5% to the enhanced greenhouse effect. In the Kyoto-conference [UNFCCC, 1997], held in December 1997, it was agreed to reduce worldwide greenhouse gas emissions of the industrialised nations by 2012 to a 5.2% lower level compared to the emissions in 1990. For the Dutch situation this agreement implies that in the period 2008-2010 compared to the 1990 level 6% less greenhouse gases will be allowed to be emitted according to the Uitvoeringsnota Klimaatbeleid[Duurzame energie in uitvoering, 1999]. This will have to be realised under higher economic grow rates and accompanied with a higher consumption level, which potentially causes increasing greenhouse gas emission levels [Energierapport, 1999]. Therefore, a reduction of the use of fossil fuels, which are the major source of CO2 emission, is required. In order to reach the goal of decreasing CO2 emissions, growing attention is drawn toward two directions. The first is to improve the efficiency of existing fossil fuel conversion processes. The second is the use of sources that are practically CO2 emission free, like wind, solar, geothermal, hydropower, biomass and advanced nuclear sources [Gardiner, 1990]. Biomass is almost CO2 neutral and renewable, because CO2 is taken up from the atmosphere during the growth of biomass and released again during combustion in a relatively short cycle time, as compared to fossil fuels. An increased use of renewable energy sources can support governments to achieve a wide range of policy goals: e.g. improved energy security and diversity, enhanced levels of technology export to countries which are less developed and reduced emissions of greenhouse gases and other pollutants, such as sulphur oxides, nitrogen oxides, particulates and trace metals. In the world, the USA is the country with the highest energy demand per capita. According to the International Energy Outlook [IEO, 2000], the share of renewables in the US electricity market is projected to be about 20% in 2020. The majority of this renewable power production will be hydroelectrity. Biomass will provide a smaller contribution to the US power supply, which is currently approximately 4% of the primary energy [Costello & Chum, 1998]. The US department of Energy (DOE) has published a figure of 1% use of grid-connected biomass power capacity, which is about 7 GW. A goal set by the US government is to implement an additional 17 GW biomass power the next two decades [DOE, 1996].2

The European Union recently published a White Paper on Renewable Energy, stating that Europe could double its use of renewable fuels from 6 % in 1997 to 12% by the year 2010. Currently, in Europe energy from biomass sources accounts for about 45 Mtoe (approximately 3% of the total consumption), while the European Commission proposes that biomass in total will contribute an additional 90 Mtoe per year by 2010 with an increased share in the total energy consumption [EC, 1997]. The Dutch government has agreed to a goal of 10% energy from renewable sources for the year 2020, with biomass as the main contributor [Weterings et al., 1999]. This implies about 120 PJ of bioenergy. Recent expectations are such that in ten years time biomass for energy supply could increase from a current level of approximately 13 PJ to a level of 80 PJ [Energierapport, 1999]. For this purpose, the Dutch government uses several instruments in a liberalising electricity market, such as offering fiscal advantages, to stimulate the demand of green electricity as well as the production of energy from renewable sources [Kwant & Leenders, 1999]. 1.2 Biomass as part of renewable power generation Since the dawn of mankind, biomass has been used as food for life and energy for heating and cooking by combustion. It is a form of solar energy stored in organic form. Nowadays it is recognised that the use of biomass for supply of energy offers the advantage of reducing the net CO2 emission. Power generating facilities based on fossil fuels, of which coal is the most important component, cause comparatively high specific emissions of the greenhouse gas CO2. Today around 40% of the worlds electricity generation is based on coal and 20% of the CO2 emissions are caused by coal-fired power plants [Campbell et al., 2000]. Biomass can nowadays be utilised in quite efficient processes for power and heat production, which have been developed to industrial scale in the last decades. For this purpose, biomass from energy crop cultivation, set-aside land can be used in an economic way in certain countries. This helps to restructure the agricultural situation with the benefit of job creation along with the agro-economic activities for this purpose. A valuable bio-diversity can be created along with erosion protection, depending on the crop used. Drought tolerance and low fertiliser and/or pesticide requirements are important requirements [Hall et al., 1993], [DOE, 1996], [Faaij, 1997]. Also, (industrial) waste streams of biological origin can be used, thereby reducing disposal problems or incineration, which is inefficient compared to electricity (co-) generation. The current commercial and non-commercial biomass use for energy production is estimated to be between 6 and 17% of world primary energy, most of this is used in developing countries, where biomass accounts for up to one third of energy needs [Bauen & Kaltschmitt, 2001]. By contrast, biomass provides at most 3% of energy in industrialised countries [Gross et al., 2003]. The utilisation of solid fossil together with biomass for heat and power production offers additional advantages compared to the use of biomass as a single fuel [Rdiger et al., 1996]:

variations in the availability of biomass can be met by changing coal-biomass ratios; it enables a wider range of system sizes, increasing optimisation possibilities; it can show synergistic effects that reduce operational problems and emissions.1.3 Technology for biomass utilisation for heat and power generation An important aspect for the choice of the technology to convert (biomass) fuels is the cycle efficiency for power and heat generation. Efficiency improvements decrease specific fuel consumption and emissions. Available energy conversion systems for solid fuels have thermodynamic limitations with respect to attainable efficiencies. Power generation in conventional systems is based on either steam turbine or gas turbine technology. The application of advanced efficient systems such as fuel cells is interesting

3

for the longer term but these are not considered in this study as they are presently comparatively expensive and not technically feasible for large scale biomass/waste based conversion processes. The first systems used at commercial scale were based on a steam turbine cycle with direct combustion of the fuel. In this conversion process heat is transferred from a high (combustion) temperature to a level determined by the maximum allowable steam conditions (pressure and temperature). These conditions are mainly determined by material strength constraints. Due to the relatively large difference in the temperatures mentioned (combustion and steam temperature), the process efficiency is limited thermodynamically by the upper temperature level of the steam cycle. A gas turbine can be used efficiently for generating power using high temperature working fluids, like flue gas from solid fuel or gasification-derived gas combustion. One of the limitations in power generating efficiency in a gas turbine (or Brayton) cycle is in the lower temperature level to which heat can be released. A combination of a gas turbine and a steam turbine, a so-called combined cycle, leads to potentially higher efficiencies as compared to the separate steam and gasturbine cycles. In the following sub paragraph, a comparison is made between small and large-scale systems for power and heat production, as these have their own advantages and disadvantages. 1.3.1 Small scale versus large scale processes Electric power from biomass can be generated in a decentral or central way. Large scale centralised power production offers the following main advantages (with specific biomass related points indicated in italic): + the availability of electricity is well secured; + the economics of scale cause comparatively low power production costs; + emissions can be controlled adequately, due to the gas cleaning equipment being relatively well known on large scale and operated by skilled personnel. Disadvantages are: users are dependent on a grid, which is sometimes not reliable in developing countries; significant losses in the distribution and transportation of electricity, although in a densely populated country, like e.g. the Netherlands, this loss effect can be ignored; transportation distances of fuel are relatively long, with accompanied increased costs and CO2 emission; the availability of especially biofuels can be insecure. In the decentralised option, which was also characteristic for traditional biomass use, electricity is produced by small scale units in local communities, or farms located far away from densely populated areas. This has the following advantages:

+ + +

users are not dependent on the electric grid, which especially in countries in the developing world is not always secured, a critical issue for e.g. hospitals; transportation and distribution losses of electricity are relatively small; the average distance of transportation of biomass fuels is short.

Disadvantages are: emissions of small local power producing equipment cannot be controlled as adequate as those of large scale power plants; maintenance of the units requires special skills and is possibly less secured as compared to large-scale power production; fuel conversion, gas cleaning (both for protection of downstream equipment and emission reduction) and prime mover equipment is relatively expensive due to the negative economics of scale;

4

availability is not always secured, as skilled personnel is not always there and also because there can be problems in biomass supply.

In the Dutch situation the advantages of large-scale systems outweigh the disadvantages and therefore the focus of this study will be on large-scale systems. 1.3.2 Technological options for large scale biomass based heat and power generation For large scale heat and electricity generation using biomass, the main technological options are combustion or gasification. Pyrolysis, or liquefaction in combination with a combined cycle e.g. is still in a conceptual stage [Siemons, 2002]. These techniques are more interesting for smaller scale decentralized heat and power production (scale-up is still under development) or for producing biofuels for transportation. 1.3.2.1 Combustion Co-combustion of biomass in large-scale utility boilers with, as additional option, co-production of electricity and heat is already a commercial common practice [Spliethoff, 2000]. The energy conversion systems based on combustion basically consist of a primary combustion section, a boiler section and a steam turbine. The combustion section ensures conversion of the chemical energy bound in the solid fuel into heat in the presence of an overall over-stoichiometric amount of air. This heat is transferred by radiation and convection to the water and the steam in the boiler and sensible heat of the flue gases; with conduction playing a minor role as heat transfer mechanism. The steam is expanded in a steam turbine to drive the generator for electricity production. Part of the energy content in the steam that cannot be converted into power, generates heat on a relatively low temperature level and can be used for e.g. district heating provided that a consumer net is commercially possible. Often the heat produced cannot be used to its full extent. The amount of power generated divided by the amount of heat produced of most biomass-fired power stations amounts to values less than 0.5 [Van den Heuvel & Stassen, 1994]. Depending on the local requirements variable amounts of heat can be utilized, but comparatively high electric power generation efficiency is the driving force in most situations. The electric efficiency of common steam cycles is determined by:

the temperature of steam at the turbine inlet; the higher the temperature (accompanied with a higher pressure) the higher the efficiency; the maximum temperature is determined by the material properties of the steam generator; the steam pressure at the turbine outlet (condenser pressure), which is determined by the temperature to which the outlet steam can be cooled down; the lower this temperature, the higher the overall efficiency; thermal losses of the boiler system; combustion efficiency; generator losses.

There are several commercial combustion system configurations available for both biomass and solid fossil fuels, with the following reactors as primary combustor types: fixed bed combustors (practically not applied anymore for solid fossil fuels), like e.g.: fixed flat or inclined grate firing units; moving flat or sloped grate ovens; combustors with bottom screw feeding; fluidised bed combustion reactors, such as: bubbling beds; circulating beds; other reactor systems, with as examples:

5

pulverised solid fuel burners (entrained flow reactors); cyclone burners. When compared to other combustion technologies fluidised beds offer the advantages of: very high flexibility with respect to fuel properties like size distribution, density, moisture and ash content; good heat transfer leading to installations with relatively small specific reactor volumes; in-situ primary gas cleaning by addition of sulphur binding compounds, like e.g. limestone, so that downstream cleaning is less extensive; relatively low reactor temperatures, leading to less corrosion, deposition problems and potentially lower NOx and SOx emissions than high temperature systems. Problems of these fluidised bed combustion systems are fouling and slagging of heat exchanging equipment in the system (boiler) and sintering tendencies of ash and bed material especially with fuels containing alkali metals and chlorine. Also, for fluidised bed combustion of coal comparatively high N2O emissions have been observed. 1.3.2.2 Gasification An alternative to direct combustion of biomass and/or coal for conversion of fuel into power and heat is conversion by thermal gasification. In principle there are three basic gasification reactor configurations: the fixed bed gasifier (dry ash or slagging), which can be distinguished into three subtypes: the co-current down flow reactor; the counter-current reactor, and the cross-current gasifier; the fluidised bed gasifier (non slagging operation), which can be divided into two sub configurations: the bubbling fluidised bed and the circulating fluidised bed; the entrained flow gasifier (slagging). The process can be configured as autothermal or allothermal. In the first, most commonly found configuration, the heat necessary to drive the endothermic gasification reactions is provided in-situ by partial combustion of the fuel. For the latter configuration, the heat needed for the endothermic reactions is provided externally. Application of oxygen/steam as gasification medium is common for coal gasification, because smaller equipment can be applied when compared to air blown gasifiers. Gas cleaning is a critical issue for the successful implementation of gasification technology for electricity generation. The combination of the gasifier, down stream followed by high temperature dry gas cleanup processes offers the advantage of higher overall efficiencies than systems using wet, low temperature, raw gas cleaning techniques. This holds especially for downstream applications that are using hot product gas. Relatively high power efficiencies can be attained by application of Integrated Gasification & Combined Cycle (IGCC) technology. High efficiency, natural gas fired combined cycles are applied on a large scale throughout the world with natural gas as fuel. These units require comparatively low investments. The IGCC process, characterised by a relatively high power to heat production, has been demonstrated already for coal as solid fuel on large scale and seems to be also especially attractive in the medium-size power production range of 20-150 MWe [Kurkela et al., 1993b]. Examples of technically successful coal-based IGCC demonstration projects in Europe are "Buggenum" (253 MWe, see figure 1.2 a and b) [Ploeg, 2000] and "Puertollano" (300 MWe) [Schellberg, 2000]. In the USA, coal-based IGCC projects are: "Polk Country" (250 MWe, 4.5 year demonstration project ended in 2001, now continuing to operate commercially), "Wabash River" (262 MWe, 5 year demonstration project ended in 1999, now continuing to operate commercially) and6

"Pion Pine" (100 MWe), see e.g. [Campbell et al., 2000]. The first two mentioned can be characterised as technically successful, with availability data as high as 70-80%; the last one has ended less successfully [Henderson, 2003]. Furthermore, also in Japan and Australia IGCC projects based on coal have been launched [Henderson, 2003].

Figure 1.2 a. Picture of the Buggenum IGCC power station; [Scheibner&Wolters, 2002].

b. Process scheme of the plant.

The IGCC process comprises a gasifier with appropriate product gas cleanup and both a gas turbine and a steam turbine for electricity generation. The competitiveness of the process alternatives has been based on the overall price of electricity, discarding the environmental benefits related to IGCC technology. In the future, the environmental factors and sustainability aspects may probably be of higher importance and competitiveness of IGCC systems based on a variety of solid fuels will then be significantly improved. However, the competitiveness of IGCC cannot be based only on environmental superiority. The overall feasibility of IGCC technology also has to be improved. This can be attained by developing and improving equipment components of IGCC units. In addition, the fuel flexibility has to be further improved one can consider all kinds of waste material in this respect - accompanied by reduction of operational costs. Fluidised bed technology is an attractive process for gasification in comparison to the fixed bed or entrained flow alternatives, because of: + the well-established principles of operation; + the fact that the process can be scaled successfully to the medium-size power range ; + its flexibility with respect to feedstock characteristics (type, particle size); + the favourable heat and mass transfer properties; + the energy-efficiency of the process?. A main disadvantage is that the fluidised bed process is limited in the temperature range to be applied, due to possibility of bed material sintering, which causes fuel conversion to be several percent points less than 100%. This disadvantage holds most for relatively unreactive fuels, such as bituminous coals, but not so for reactive biomass and brown coal. The most attractive gasification medium in the fluidised bed gasification process is air since no expensive air separation unit is necessary. A pressurised gasification process has as a major advantage that compression of the product gas, necessary for combustion in (advanced) gas turbines, is not required to such an extent as for atmospheric gasification. Compression of air for the gasifier is in this sense cheaper and easier to accomplish than compression of the fuel gas. Finally, reduced size and, hence, less costly equipment is sufficient for the gasifier and gas cleanup section as compared to atmospheric gasification processes.

7

As pressure resistant materials are more costly this argument holds for scales in a range from 20-30 MW electric and higher. As an example the process scheme of the 18 MWth pressurised fluidised bed IGCC demonstration plant at Vrnamo is given in figure 1.3. The commercialisation of combined cycle power generation based on the gasification of solid biomass fuels has been slower than expected about one decade ago. An overview of these processes is given in chapter 2.

Figure 1.3 Process scheme of the Vrnamo pressurised fluidised bed IGCC demonstration unit [Sthl, 2001] 1.4 Power production from biomass gasification, open research questions There are still partly unanswered questions, which have to be answered by means of measurement and process modelling. These questions are related to the fate of species harmful to equipment and environment based on biomass gasification with fluidised beds as primary solid fuel converters. They are summarised below:

What is the mechanism of the formation and destruction of higher hydrocarbon species (tars),which can clog process equipment parts like fuel control valves and gas analysis equipment and can cause a decreased downstream gas turbine combustion efficiency and soot emission? What is the fate of trace elements; will they be emitted to the environment in solid form as ashes (from bed and fly ash) or in gaseous form (air toxics) in the exhaust gas? How is fuel bound sulphur partitioned in the gasifier into mainly H2S and COS and subsequently SOx during gas turbine combustion? What is the fate of nitrogen species (like NH3 and HCN) in NOx formation in processes including gas turbine combustors downstream of the gasifier, when dry, high temperature gas cleaning processes are applied [Hoppesteyn, 1999]? The first open research question is not dealt with in this thesis. The second is the subject of another thesis from our institute [nal, 2005]. Regarding the remaining questions a further consideration is given below. Table 1.1 gives an overview of typical elementary compositions of a selection of coal and biomass species. The biomass examples are from agricultural waste origin (straw types), energy crop cultivation (miscanthus) and forestry.

8

Coals presented in the table below are brown coal (from the German Hambach open mining) and black bituminous coals from the UK (Daw Mill coal) and the USA (Pittsburgh and Utah coal). Data for peat (from eastern Finland), a fuel intermediate between biomass and coal, is also given. Table 1.1 Fuel composition of several solid fuel sources (on dry basis) related to their nitrogen and sulphur species emission potential.Biomass Rice straw 1) Miscanthus 2) Sawdust 3) Wood pellets (clean)2) Peat and Coal Finnish peat 4) Hambach Brown coal 2) Daw Mill coal Pittsburgh coal 1) Utah coal 1)1)(mass%, dry)

C

(mass%, dry)

H

(mass%, dry)

O

(mass%, dry)

N

(mass%, dry)

S

(MJ/kg)

HHV 15.4 19.3 19.2 20.3 21.8 25.2 30.0 31.8 33.0

(kg/GJ)

N

(kg/GJ)

S

39.2 48.2 48.5 51.4 54.5 64.5 69.4 75.5 77.9

5.1 5.4 5.1 6.0 5.6 4.4 4.4 5.0 6.0

35.8 42.8 46.0 42.2 33.6 26.1 10.0 4.9 9.9

0.6 0.58 0.03 0.17 1.80 0.64 1.2 1.2 1.5

0.1 0.16 0.03 0.11 0.25 0.39 1.6 3.1 0.6

0.4 0.3 0.02 0.08 0.8 0.3 0.6 0.4 0.5

0.07 0.08 0.02 0.05 0.1 0.2 0.8 1.0 0.2

[Reed,1981] 2) this study 3) [Zhou,1998] 4) [Kurkela et al., 1992]

Coal is a sedimentary rock composed of both organic and inorganic constituents and formed through partial decomposition of plant debris under the action of heat, pressure and time, [Bend, 1992], [Van Krevelen, 1993]. The main difference between coal and biomass is the coals higher heating value, which is closely related to the oxygen content in the fuels. During carbonisation, oxygen and hydrogen are reduced and coals of increasing rank from lignite to anthracite are formed, see [Speight, 1983]. Here, coal rank is defined as an indicator for the stage of alteration, or degree of coalification, attained by a particular coal. The greater the alteration, the higher the rank of the coal. Related to this difference in elementary composition is the much higher reactivity of biomass as compared to coal in thermochemical conversion processes, because less stable oxygen containing structures are present in the fuel. Absolute sulphur content is higher in the coals presented than in biomass types shown. This is also the case, although somewhat less pronounced, when the sulphur contents are compared on an energy basis. As sulphur species emission for biomass gasification is not a major problem for a broad range of these fuels, this issue is not further studied in this context. Although the absolute values of the nitrogen content on mass basis in biomass materials in most cases is significantly lower than coal, the nitrogen quantity on an energy basis is comparable for certain biomass species, like miscanthus and straw. This indicates the relatively high emission potential of significant nitrogen containing species, like NH3 and HCN as NOx precursors. The chemical structures in which the nitrogen is bound, however, are of importance too for the still not completely understood partitioning behaviour. This is the reason why the fate of nitrogen species will be the focus of this thesis. 1.5 The fate of fuel bound nitrogen The emission of NOx, a collective noun for NO and NO2, is a worldwide occurring regional problem as its effects are both on the environment (plants, buildings) as well as human and animal health. One of its effects is contribution to photochemical smog formation, in which NOx, CH3COO2NO2 (peroxylacetyl nitrate, PAN) and O3 play a role [Leighton, 1961]. Smog affects health negatively and reduces visibility. NOx also participates in the formation of O3 in the troposphere causing problems for vegetation. It is also a contributing agent in the greenhouse effect. Nitrous oxide, N2O, is known as laughing gas. When it is inhaled, it can give uncontrollable laughter, further inhalations lead to enjoyable hallucinations and longer exposure leads to anaesthesia [Hayhurst & Lawrence, 1992]. N2O has a relatively long lifetime in the troposphere of approximately 150 - 170 years [Badr & Probert,

9

1992] and [Hayhurst & Lawrence, 1992]. It can therefore be transported to the higher regions of the stratosphere, where it is destroyed by photolysis and a reaction with O atoms to NO as important product. NOx herein acts as a catalyst in stratospheric ozone destruction via these reactions [Badr & Probert, 1993a]. Finally, NOx contributes also to eutrophication and to acid rain or deposition, which negatively influences plant growth, aquatic life and fertile soil, as well as promotes corrosion and erosion of man-made constructions. It is estimated that NOx contributes about 30% of the acidity of rain while SO2 accounts for the rest [Badr & Probert, 1993b]. The main source for NOx in the environment nowadays is the combustion of fossil fuels in stationary (power plants, industry, and to a lesser extent households) and mobile (traffic) sources. Natural sources are the formation from living microbiological species and lightning. NO formed in combustion processes is generally discharged to the atmosphere and subsequently converted to NO2. The air oxidation of NO to NO2 at room temperature in is a relatively slow process. N2O emissions by human activity arise from chemical industry (e.g. nitric acid production). Also, fluidised bed combustion of higher rank coals cause significant N2O emissions. When air blown gasification of biomass and/or coal is applied as thermal conversion technology, a product gas is generated with heating values in the range of 3-6 MJ/kg, which contains relatively high concentrations of fuel-bound NOx precursor species [Hoppesteyn, 1999]. These are NH3, HCN and nitrogen-containing hydrocarbons, as has been measured by several investigators, e.g. [Chen, 1998], [Hmlinen, 1996], [Kurkela et al., 1996], [Kurkela et al., 1995], [Kurkela et al., 1993b], [Kurkela and Sthlberg, 1992], [Zhou, 1998]. The NOx emission from gas turbines integrated in an IGCC process is mainly formed from these fuel-nitrogen species. The formation of thermal and prompt NOx during combustion of LCV gas is negligible compared to above-mentioned fuel nitrogen derived NOx [Hoppesteyn, 1999]. N2O emissions have not been reported from gas turbine combustion processes and are not expected due to the relatively high combustion temperature range. There are several potential ways to reduce NOx emissions:

use of advanced low-NOx burners in the combustion stage; conventional gas cleaning (wet scrubbing); cleanup of the flue gas of gas turbine combustors with e.g. selective catalytic NOx reduction cleaning of the gas produced by the gasifier by e.g. selective catalytic N-species reduction (SCR), use of fuel with low nitrogen content; increase the conversion of NH3 and HCN into N2 in the primary gasification process.or selective catalytic oxidation (SCO); (SCR);

The use of low fuel-NOx burners for gas turbines is an attractive option, which is also still subject of study, see e.g. [Hoppesteyn, 1999], [Nakata, 1996].

Conventional gas cleaning, or wet scrubbing, to remove NH3 from the product gas, is an option but this has disadvantages like lower overall power generation efficiencies, as substantial product gas cooling is necessary; see [Woudstra & Woudstra, 1995] for an exergy analysis of IGCC systems with wet and dry hot gas cleaning. Also a large spectrum of organic (tars) and inorganic chemical species (like e.g. NH3, HCN, H2S etc.) has to be removed from the scrubber water before it can be recycled or released to a water treatment facility. The use of an end-of-pipe solution such as catalytic NOx removal downstream of the gas turbine, e.g. SCR (selective catalytic removal by NH3 injection), is limited due to catalyst poisoning by the product gas, particulates or temperature limitations. This solution is also less attractive because of the large gas flows, which have to be handled due to the high amount of excess air added to the gas turbine, which requires relatively big equipment. When SCR is applied under non-optimised process conditions, some of the injected, unreacted NH3 is vented to the atmosphere, thus generating a new and not acceptable pollution problem. Application of SCR upstream of the combustor by means of injection of NO, necessary to reduce NH3, can lead to catalyst poisoning by sulphur species, heavy metals and solid deposits.10

This imposes additional requirements on the gas cleaning system and this process is not yet applied at commercial scale. The option of using low nitrogen containing fuels is not always available. When agricultural residues or waste material from biological origin are applied, there is an amount of nitrogen in the fuel on an energy basis comparable to fossil fuels, as illustrated by the figures in table 1.2. Wood seems to be a good alternative in this respect, though. An increase in the conversion of fuel bound nitrogen into harmless molecular nitrogen upstream of the combustion and gas cleaning section is an attractive and promising option. The fuel bound nitrogen species concentrations in the product gas of the gasifier are decreased by measures taken within the gasifier, which leads to less extensive and thus less expensive gas cleaning before the gas turbine combustion. For the high temperature, dry gas cleanup, the argument is quite clear. With respect to wet scrubbing using water, it is necessary to emphasize that the bound nitrogen species (especially NH3) have to be removed from the scrubber water. Therefore acid chemicals are needed to neutralise the water or the ammonia has to be removed by stripping processes. The use of neutralizing agents contributes negatively to the sustainability of the biomass gasification process, due to high consumption of chemicals that are produced using fossil fuels. The ammonia stripping contributes to lower overall process efficiencies because the concentrated ammonia needs to be dealt with. Therefore, selective primary conversion of bound nitrogen into N2 in the gasifier seems to be the better solution. However, there is still no clear picture how the fuel bound nitrogen in biomass is converted into NH3, HCN, nitrogen-containing hydrocarbons, nitrogen oxides and N2 during the gasification process. In particular for pressurised (fluidised bed) gasification no model including bound nitrogen conversion has been set up and tested against experimental data so far. There are also no extensive concentration profile measurements of nitrogen species before and downstream of high temperature gas cleaning units available in the open literature. This is partly due to measurement difficulties and partly caused by confidentiality policies of gasifier manufacturers. The aim of this thesis is to set up a reactor model that is capable of predicting the main product gas composition and fuel bound nitrogen speciation using experimental data from fuel characterisation experiments and validate it with measurements carried out at pilot scale (50 - 1500 kWthermal) pressurised air-blown fluidised bed reactors. 1.6 Outline of this thesis

This thesis deals with experimental and modelling work concerning biomass/coal pressurised airblown bubbling fluidised bed gasification. Emphasis is put on the formation and destruction of nitrogen compounds from solid fuel bound nitrogen, which to the knowledge of the author has not been considered in models for such gasifiers so far. The present study is limited to steady state performance. Start-up, shutdown or other dynamic phenomena are not studied. The influence of the fuel sources and additives as well as process conditions like pressure, temperature and reaction stoichiometry on the main gas constituents and gaseous nitrogen species are investigated. The experimental validation of the developed model is performed on two scales of thermal input. In chapter 2 an overview is given of biomass based commercial fluidised bed gasifiers and also of experimental fluidised bed gasification research and development conducted throughout the world. Furthermore, special attention is paid to literature concerning modelling of the fundamental stages of fluidised bed gasification and to experiments performed to determine the fate of nitrogen species in gasification processes. Potentially attractive primary measures for fuel-NOx emission reduction are highlighted. Finally, the experimental and theoretical objectives of this study are presented. Chapter 3 describes the pressurised fluidised bed test rigs of Delft University of Technology and Stuttgart University that have been used for the experimental validation experiments at two scales. Furthermore, a TG-FTIR set up is presented which has been used for the characterisation of the pyrolysis process of fuels used in the gasification experiments.

11

Finally the pressurised heated grid reactor at Eindhoven University of Technology is described, which has been used for flash pyrolysis experiments of fuels used in the experimental programme of this research. Chapter 4 presents the results of the gasification tests conducted in both pressurised fluidised bed test installations. Also, pyrolysis experiments performed with the TG-FTIR set-up are presented and discussed in relation with the gasification experiments. Finally experimental results from heated grid flash pyrolysis are presented and analysed. Chapter 5 deals with the modelling of the pressurised fluidised bed gasification process with emphasis on the fate of fuel bound nitrogen species. Here, simulation cases are presented with the aim to test the model for various process conditions and to study the expected effect of primary measures to reduce fuel bound nitrogen content in the gasification product gas. In chapter 6, modelling and experimental results for the pressurised fluidised bed tests carried out at two scales of thermal input are compared and analysed. Finally, in chapter 7 conclusions of the experimental work, the modelling study, the comparison of model and experiments are summarised. In addition, recommendations for further research are presented.

12

Chapter 2

Fluidised bed solid fuel gasification processes, overview and analysis of experimental research and modelling2.1 The fluidised bed reactor applied for solid fuel gasification

The fluidised bed reactor has been one of the workhorses for large scale coal conversion, since it was first patented by Fritz Winkler in 1922 and commercialised in 1926, see e.g. [Howard, 1983] and [Kunii & Levenspiel, 1991]. The reactor scaling up from bench scale units via pilot plants to large units for combustion or gasification has been mostly carried out empirically. Emission studies from fluidised bed thermal conversion processes have often also been empirical. Modelling studies, however, concerning fuel conversion and main product gas components generation for gasification have been ongoing since a few decades. In this chapter an overview is given of biomass based commercial fluidised bed gasifiers. Hereafter recent experimental fluidised bed gasification research and development conducted worldwide both on small (5 MWth) fluidised bed biomass / waste gasification projects, an overview classified according to thermal input scale.ProjectFICFB process Gssing (Austria) BioCoComb, Zeltweg (Austria) Skive Fjernvarme (Denmark) Enviropower, Tampere (Finland) Biomass Gasifier Facility (BGF), Hawaii (USA) Bioflow Vrnamo (Sweden) ARBRE Aire Valley (United Kingdom) Aomori Automotive Shredder Plant (Japan) Battelle process (FERCO), Burlington (Vermont, USA) HTW Rheinbraun, Wesseling (Germany) Corenso Varkaus (Finland) THERMIE Energy Farm, Pisa (Italy) Electrabel Ruien (Belgium) Kymijrvi power plant Lahti (Finland) Amer power station (Essent), unit 9, Geertruidenberg (NL) Aerimpianti, Greve-in-Chianti (Italy) BIG-GIT Bahia (Brasil) HTW Rheinbraun, Berrenrath (Germany)

GasifierAustrian Energy & TU Wien, CFB Austrian Energy & Environment CFB Carbona Pressurised BFB Carbona Pressurised BFB IGT RENUGAS Pressurised BFB Foster Wheeler Pressurised CFB TPS Atmospheric CFB Ebara corporation ABFB Battelle Columbus interconnected CFBs Pressurised BFB, Lurgi licensed Foster Wheeler/ Corenso Ltd. Atmospheric BFB Lurgi Atmospheric CFB Foster Wheeler Atmospheric CFB Foster Wheeler Atmospheric CFB Lurgi Atmospheric CFB TPS Atmospheric CFB TPS Atmospheric CFB Pressurised BFB, Lurgi licensed

Use of product gasGas engine 4.5MWth, 2 MWel Co-firing in pulverised coal combustor, 10 MWth biomass / 330 MWth coal BIG gas engine, 11.5 MWth, 5.5 MWe 15 MWth Gasification testing for electricity and methanol production (50/50), 90 tonne/day (15 MWth,max) IGCC, 18 MWth, 6 MWe IGCC, 8 MWe Boiler for power generation 2 x 40 MW Initially in steam cycle power plant, later in IGCC 15 MWe Demonstration unit for IGCC, approx. 35 MWth Use of gas in boiler 40 MWth IGCC, 16 MWe Co-firing in Pulverised coal/natural gas boiler, 50 MWth Co-firing in pulverised coal/natural gas boiler, 40-70 MWth Co-firing in Pulverised coal combustor, 80 MWth Combustion of LCV gas, power generation in steam cycle IGCC, 32 MWe Demonstration unit for methanol production, 140 MWth

Current statusIn operation In operation from 1997-2001, shut down after successful operation Contract signed, construction fall 2004 Shut down after demonstration Shut down after demonstration in 1998 Shut down after successful demonstration, 1999 Start-up phase ended, now mothballed In operation

References[Hofbauer et al., 2003] [Anderl et al., 1998] [Fernando, 2002] [Patel, 2004] [Salo & Patel, 1997] [Lau, 1998], [Wiant et al., 1998], [Lau & Carty, 1994] [Sthl, 1998] [Sthl, 2001] [Pitcher et al., 1998] [Pitcher et al., 2002] [Selinger et al., 2003]

Demonstration Shut down after demonstration in period 1989-1992 In operation Construction procedure pending Under construction In operation

[Paisley et al., 1997] [Adlhoch et al., 1992a,b] [Opet, 2001] [DeLange & Barbucci, 1998] [Anttikoski, 2002] [Raskin et al., 2001], [Nieminen & Kivel, 1998] [Willeboer, 1998] [Greil et al., 2000] [Fernando, 2002] [Rensfelt, 1997] [Barbucci, 1999] [Arrieta & Sanchez, 1999] [Maniatis et al., 2003] [Adlhoch et al., 1992a]

Start-up phase

In operation Definition, progress unclear Shut down after demonstration in period 1986 - 1997

2.3

An overview of recent research, development and small scale demonstration activities

In the past two decades quite a number of fluidised bed gasifiers for biomass/coal research and process development have been operated in thermal input range up to 3 MW. Table 2.2 presents an overview of the process development installations having a solid fuel input capacity of 50 kWthermal and higher, which have been relatively well documented in the literature.

17

Table 2.2 Overview of fluidised bed biomass (+ solid fossil fuel) gasification process development test rigs (>50 kWth) and small scale (semi-) commercial plants (50% of particles < 0.28 mm). Another explanation of the observed increase in fuel-N conversion to NH3 with temperature given by the authors was the lower char content with higher temperatures, leading to a decreased char-NO reduction to N2. A different trend was observed in a bench scale ABFB gasifier by [Zhou, 1998]. Increasing the temperature, while keeping the air stoichiometry the same, lead to a drastic decrease in the fuel-N to NH3 conversion and an increase in the conversion to N2 for several biomass fuels in an atmospheric ABFB. For Leucaena, a high N containing fuel, fuel-N conversion to NH3 decreased from about 64 to 18% with temperature increasing in the range of 750C to 950 C. For HCN the values decreased from 0.10 to 0.07%. Tar-N was not observed, though measured, and char-N decreased from 7.7% to 1.2%. In the same range fuel-N to N2 increased from about 39 to 86%. In his experiments the N balance closure was between 105 and 116%. In general the observations are such that fuel nitrogen conversions to NH3 and HCN can be expected to decrease with temperature. The exceptions reported by [Zhang et al., 2002] and [Rosn et al., 1997] could be attributed to low fuel conversions, so that at increasing temperature the solid fuel conversion increases and thereby the conversion to NH3 as well. 2.4.4 Influence of pressure The Finnish VTT group investigated the effect of pressure during bottom-fed fluidized bed gasification of peat and wood. The results indicate that increased pressure favours the formation of NH3 and the NH3/HCN ratio is increased (see [Lepplahti & Kurkela, 1991]). On the other hand, researchers at KTH found a slightly decreasing trend of the fuel nitrogen conversion to NH3 with increasing pressure in their top-fed fluidised bed gasifier for birch gasification [Rosn et al., 1997].

25

[Nichols et al., 1987] found that the NH3 concentration increased, whereas the HCN concentration decreased with increasing pressure in an entrained flow coal gasifier. [Zand et al., 1985] reported that pressure had no measurable effect on the production rates of the nitrogenous compounds for fluidised bed peat and lignite gasification, although their applied pressure range was limited (770-830 kPa). The reported influence of pressure on fuel nitrogen partitioning is such that at this stage no conclusive statements can be made. In this respect, the opposing observations by KTH and VTT can possibly be related to their different feed location configurations; KTH operated a top fed gasifier, versus VTT and most of the mentioned references a bottom fed gasifier. 2.4.5 Influence of additives Additives in the fluidised bed gasification of solid fuels are used for several purposes: reducing gaseous sulphur emissions, effective for relatively high sulphur containing fuels, see e.g. [Thambimuthu, 1993] and [Kurkela et al., 1996]; cracking of tars, see e.g. [Milne et al., 1998]; prevention of agglomeration, see for instance [Moilanen & Kurkela, 1998] and [Turn et al, 1998b]; catalysis of the process for e.g. hydrogen yield optimisation, see e.g. [Rapagna et al., 2000 & 1998]. Especially for the first three purposes, dolomite or limestone find widespread application. The use of (calcined) dolomite and/or limestone as additive in FB gasification resulted in decreasing HCN and increasing NH3 concentrations, see e.g. [Lepplahti & Kurkela, 1991]; [Lepplahti & Koljonen, 1995]. This effect was reported in literature, where CaO was suggested to react with HCN, forming a relatively instable CaCN2 compound, see [Lepplahti et al., 1991]. The reaction of CaCN2 with water (vapour) back to CaO, forming CO and NH3, is thermodynamically favoured, see [Jensen, 1996]. Also, [Berg et al., 2001] cited confidential reports on biomass gasification ([Waldheim et al., 1999] and [Blackadder et al., 1995]), showing that dolomite and limestone addition favoured NH3 formation at the expense of HCN. [Paterson et al., 2002] experimentally found that in their pressurised spouted bed gasifier addition of limestone to Daw Mill Coal led to increased NH3 concentration in the exit gas. Iron in the fuel was reported to decrease the fuel nitrogen conversion into NH3 significantly, see [Lepplahti, 1998] and [Lepplahti et al., 1991]. 2.4.6 Influence of particle diameter From their CFB gasification observations using differently sized wood particles as feedstock, [Van der Drift & Van Doorn, 2001] report that sawdust with a maximum diameter of 2 mm shows lower fuel nitrogen conversions into NH3 than increasingly larger cylindric willow wood particles of 10x10 and 10x40 mm. When small sized sawdust was mixed with the larger cilindric wood particles lower NH3 conversion values were observed. The authors did not explain this effect for the fuel nitrogen behaviour, but reported that the riser showed significantly different temperature profiles. With fine sawdust large temperature differences are created over the reactor, accompanied with relatively high maximum temperatures, whereas with the larger particles a more even temperature distribution was observed with lower maximum temperature. Possibly the higher temperatures obtained by feeding fine sawdust have contributed to NH3 destruction.

26

2.4.7 Influence of steam The use of steam prevents peak temperatures that can lead to ash sintering and melting. Steamflow is therefore a process parameter used to avoid operational problems caused by sintering [Moilanen&Kurkela, 1998]. Addition of steam in air/N2 coal gasification of Daw Mill Coal in a small scale spouted bed reactor at a pressure of 1.3 MPa and a temperature range of 770 (upper bed)-900 C (spout) lead to almost complete conversion of the fuel bound nitrogen into NH3, see [Paterson et al., 2002] and [Zhuo et al., 2002]. Significantly lower fuel-N conversions into NH3 were reported when no steam was used as (almost a factor 10 lower). The authors attributed the observed effect to the reaction of steam or H2, produced by steam gasification, with char bound nitrogen. The observation confirms experimental results with steam gasification of a range of coals in a combined fluidised bed fixed bed reactor system by [Chang et al., 2003]. 2.4.8 Influence of feed location At KTH (Stockholm, Sweden) fluidised bed biomass gasification research is carried out since about two decades. Pressurised gasification is performed in a top-fed 80 kWth process demonstration unit. Relatively low conversion values of fuel bound nitrogen into NH3 and HCN are reported as compared to other fluidised bed units, which are basically all bottom fed reactors. For birch wood gasification at 0.4 MPa and approximately 900 C, fuel-N conversion to NH3 amounted to 12-21%. For conversion into HCN these values were in the range 0-0.7% and virtually no char-N was left. For experiments at the same pressure, but at temperatures of appoximately 700 C, the fuel-nitrogen conversion to NH3 was as low as 0-1.6%, conversion into HCN was 0-0.1%, and char-N amounted to 2.6-13.6%. NO formation was observed in the range 0-11.1%. The particle size of the wood feedst