You are on page 1of 177

Diss. ETH No.

17741

SOLAR GASIFICATION OF
CARBONACEOUS MATERIALS

REACTOR DESIGN, MODELING
AND EXPERIMENTATION

A dissertation submitted to
ETH ZURICH

for the degree of


Doctor of Sciences

presented by
ANDREAS Z’GRAGGEN
Dipl. Masch.-Ing. ETH
born July 25, 1978
citizen of Schattdorf (UR)

accepted on the recommendation of


Prof. Dr. Aldo Steinfeld, examiner
Prof. Dr. Michael F. Modest, co-examiner

2008
Abstract

Solar steam-gasification of carbonaceous materials is proposed as an intermediate


step on the path toward a sustainable energy economy. The developed reactor
technology, together with a set of experimental results, is presented in the first part
of this thesis. Numerical models, described in the second part, were used to support
the engineering design in the first place and to gain further insight in the heat and
mass transfer processes affecting the reactor’s performance.
The developed reactor features a continuous vortex flow of steam laden with
feedstock particles confined to a cavity-receiver and directly exposed to concen-
trated solar radiation. This setup provides efficient radiative heat transfer to the
reaction site to drive the high-temperature highly endothermic process. A 5 kW pro-
totype reactor was tested in a high-flux solar furnace with three distinct feedstocks:
(1) Dry coke powder with steam fed separately yielded up to 87% coke conversion
in a single pass of 1 s residence time at temperatures in the range 1500–1800 K.
The solar-to-chemical energy conversion efficiency was 9%. (2) Coke-water slurry
injected continuously resulted in coke conversion of up to 87% and energy conver-
sion efficiency of up to 5%. Finally, (3) liquefied vacuum residue with steam fed
coaxially yielded maximal coke conversion of 50%, and energy conversion efficiencies
of 2%. The inferior efficiencies of the latter two feedstocks were either due to excess
water used to produce the slurry (2), or to particle deposition due to non-optimal
vacuum residue injection (3). Typical syngas composition produced was 60% H2 ,
26% CO, 12% CO2 , and 2% CH4 for delayed coke and 71% H2 , 16% CO, 9% CO2
and 4% CH4 for vacuum residue. The results indicate the technical feasibility of
simultaneous pyrolysis and steam-gasification of delayed coke particles in the range
2–200 µm and of liquefied vacuum residue using concentrated solar energy.
A lumped-parameters steady-state process model that couples radiative heat
transfer with the reaction kinetics is used to support the engineering design. In
a subsequent more detailed simulation, the reactor is modeled by means of a two-
ii Abstract

phase formulation that couples radiative, convective, and conductive heat transfer
to the chemical kinetics for polydisperse suspensions of reacting particles. The gov-
erning mass and energy conservation equations are solved by applying advanced
Monte-Carlo and finite-volume techniques. Validation is accomplished by compar-
ing the numerically calculated temperatures, product compositions, and chemical
conversions with the experimentally measured values obtained from testing a 5 kW
prototype reactor.
The validated reactor model is then used to optimize the reactor’s geometrical
configuration and operational parameters (feedstock’s initial particle size, feeding
rates, and solar power input) for maximum reaction extent and solar-to-chemical
energy conversion efficiency of a 5 kW prototype reactor and its scale-up to 300 kW.
The advantageous volume-to-surface ratio of the 300 kW scaled-up reactor and its
enhanced insulation lead to a solar-to-chemical energy conversion efficiency of 24%.
Moreover, the increased cavity dimensions permit the use of a coarser feedstock to
be reacted efficiently. Finally, the tube-shaped cavity was found to outperform the
commonly employed barrel-shaped cavity, mainly because most of the particles are
directly exposed to the incoming high-flux solar radiation.
Riassunto

Il processo di gassificazione solare di materiali ricchi in carbonio, usando vapore


acqueo, è proposto come passo intermedio sulla strada verso una economia basata
sull’uso sostenibile di energia. Nella prima parte di questa tesi viene descritto il
reattore prototipo e la rispettiva tecnologia sviluppata, seguiti da risultati speri-
mentali ottenuti per diversi edotti. La seconda parte della tesi è incentrata sullo
sviluppo di modelli numerici, i quali da un lato sono stati utilizzati come base per la
progettazione ingegneristica, dall’altro hanno permesso una migliore comprensione
dei processi di scambio di calore e massa che influenzano le prestazioni del reattore.
Nel reattore presentato le particelle di materia prima sono sospese in un flusso
continuo di vapore acqueo a forma di vortice, confinate in una cavità e direttamente
esposte a radiazione solare concentrata. Questa configurazione permette il trasfer-
imento efficiente di radiazione calorica agli edotti coinvolti in una reazione chimica
altamente endotermica che procede ad alte temperature. Un prototipo del reattore,
operante a 5 kW, è stato testato in un forno solare a flusso elevato con tre edotti
distinti. (1) Polvere asciutta di coke con vapore acqueo immesso separatamente è
risultata in conversioni di coke fino a 87% con un tempo di reazione medio di 1 sec-
ondo e temperature tra i 1500 e 1800 K. L’efficienza della conversione energetica da
solare a chimica ha raggiunto un massimo di 9%. (2) Un miscuglio di coke e acqua,
chiamato slurry, iniettato a getto continuo è risultato in conversioni di coke fino a
87% ed efficienze fino a 5%. Infine, (3) il residuo di vuoto (vacuum residue) liquido
con vapore acqueo iniettato coassialmente è risultato in conversioni fino a 50% ed
efficienze fino a 2%. L’efficienza più debole degli ultimi due edotti usati è dovuta
all’eccesso di acqua necessario per la produzione dello slurry (2), o al deposito di
particelle per l’iniezione non ottimale del residuo di vuoto (3). La composizione tipi-
ca del gas di sintesi (syngas) prodotto è stato di 60% H2 , 26% CO, 12% CO2 , e 2%
CH4 per le particelle di coke e di 71% H2 , 16% CO, 9% CO2 e 4% CH4 per il residuo
di vuoto. I risultati indicano la fattibilità tecnica della pirolisi e gassificazione a
iv Riassunto

vapore acqueo simultanea di particelle di coke con diametri tra i 2 e i 200 µm e di


residui di vuoto liquidi, utilizzando energia solare concentrata.
Per appoggiare la progettazione ingegneristica è stato utilizzato un modello
basato su parametri medi che considera lo stato stazionario del sistema e che collega
i processi di trasmissione della radiazione calorica alla cinetica della reazione chim-
ica. In una successiva simulazione più dettagliata, il reattore è stato modellato per
mezzo di un flusso a due fasi che considera la trasmissione di calore per radiazione,
convezione, conduzione e la cinetica chimica per una sospensione di particelle di varie
dimensioni. Le equazioni di conservazione di massa ed energia sono state risolte me-
diante l’applicazione di tecniche Monte-Carlo e mediante il metodo dei volumi finiti.
La validazione è stata eseguita confrontando risultati calcolati numericamente per le
temperature, la composizione del gas prodotto e la conversione chimica con i valori
ottenuti in via sperimentale dai test con il reattore prototipo.
Il modello validato è poi stato utilizzato per ottimizzare le proporzioni geomet-
riche del reattore e i parametri operativi (dimensione delle particelle, rata d’immissione
degli edotti, energia solare incidente) per l’ottenimento di una conversione chimica
oppure di un’efficienza energetica massima. Sono stati considerati due reattori: da
un lato il prototipo testato a 5 kW e dall’altro un reattore simile progettato per un
input solare di 300 kW. Il rapporto vantaggioso volume a superficie del reattore a
300 kW e l’isolazione termica migliore sono risultati in un’efficienza energetica di
24%. Inoltre, l’aumento delle dimensioni della cavità consentono l’impiego di par-
ticelle di dimensioni più grandi. Infine, una cavità a forma di tubo è risultata più
adatta delle cavità a forma di barile comunemente impiegate, soprattutto perché la
maggior parte delle particelle è in tal modo direttamente esposta a radiazione solare
altamente concentrata.
Acknowledgments

First of all, I thank Prof. Dr. Aldo Steinfeld for supervising my doctoral studies at
the Professorship in Renewable Energy Carriers, ETH Zurich. The excellent research
environment he provided, his confidence in my independent way of working as well
as his lasting support were essential to my work.
I am very grateful to Prof. Dr. Michael F. Modest for critically reviewing this
manuscript and for acting as co-examiner.
The members of the Synpet project team deserve special acknowledgment. I
thank Dominic Trommer and Philipp Haueter for giving me insight in the fields of
chemical engineering and reactor design, respectively. I will keep in good memory
the weeks we spent struggling with our experimental setup. I am also indebted to
the technical staff at PSI for their support, they made the experimental campaigns
possible in the first place.
I thank all my colleagues and members of our research group. In particular,
my thanks go to Dr. Wojciech Lipiǹski and Dr. Jörg Petrasch for many fruitful
discussions on radiative transfer and numerical simulations. Hansmartin Friess
also deserves special thanks for his patient support in the development of the
gas-temperature measurement apparatus. I thank Dr. Thomas Osinga, Dr. Elena
Gálvez, Dr. Peter von Zedtwitz, Tom Melchior, Gilles Maag, Viktoria von Zedtwitz-
Nikulshyna, Patrick Coray, Sophia Haussener, Nic Piatkowski and Lothar Schunk
for providing an inspiring working environment.
Last, but no least, my very special thanks go to those close to my heart. My
parents Erwin and Susanna for igniting my curiosity and for giving me the belief
that everything is possible. Fabia for the daily royal treatment.
vi Acknowledgments
Contents

Abstract i

Riassunto iii

Acknowledgments v

Nomenclature xi

1 Introduction 1
1.1 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Conventional Gasification . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Solar Thermochemistry and Solar Gasification in Particular . . . . . . 6

I Experimental Work 9

2 Feedstock 11
2.1 Petrozuata Delayed Coke . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Characterization of Coke-Water Slurries . . . . . . . . . . . . 16
2.2 Petrozuata Vacuum Residue . . . . . . . . . . . . . . . . . . . . . . . 18

3 Experimental Setup 19
3.1 Reactor Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2 Feedstock Injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2.1 Dry Powder Feeding . . . . . . . . . . . . . . . . . . . . . . . 22
3.2.2 Slurry Feeding . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.3 Liquefied Vacuum Residue . . . . . . . . . . . . . . . . . . . . 24
3.3 Peripherals and Facility . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.4 Key Process Parameters . . . . . . . . . . . . . . . . . . . . . . . . . 27
viii Contents

3.4.1 Product Gas Composition . . . . . . . . . . . . . . . . . . . . 27


3.4.2 Chemical Conversion . . . . . . . . . . . . . . . . . . . . . . . 28
3.4.3 Energy Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.4.4 Residence Time . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4 Results 33
4.1 Coke — Dry Powder Feeding . . . . . . . . . . . . . . . . . . . . . . 33
4.2 Coke — Slurry Feeding . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Vacuum Residue — Liquefied Injection . . . . . . . . . . . . . . . . . 42
4.4 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

II Modeling 49

5 Simulation Framework 51
5.1 General Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . 51
5.1.1 Incoming Radiation . . . . . . . . . . . . . . . . . . . . . . . . 51
5.1.2 Reactor Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.1.3 Quartz Window . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.2 Polydisperse Coke Particles . . . . . . . . . . . . . . . . . . . . . . . 58
5.2.1 Radiative Properties . . . . . . . . . . . . . . . . . . . . . . . 60
5.3 Chemical Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

6 System Analysis — Lumped Parameters Model 67


6.1 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

7 Heat and Mass Transfer in the Reactor Cavity 77


7.1 Heat Transfer Modes in the Polydisperse Particle Suspension . . . . . 78
7.2 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.3 Numerical Implementation . . . . . . . . . . . . . . . . . . . . . . . . 82
7.3.1 Monte-Carlo Ray Tracer . . . . . . . . . . . . . . . . . . . . . 84
7.3.2 Fluid Flow Solver . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.3.3 CFD-MC Coupling . . . . . . . . . . . . . . . . . . . . . . . . 91
7.4 Results and Validation . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Contents ix

8 Optimization and Scale-up 105


8.1 Simulation Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
8.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

9 Conclusions 115
9.1 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
9.2 Heat and Mass Transfer Modeling . . . . . . . . . . . . . . . . . . . . 116
9.3 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

III Appendices 119

A Gas Temperature Measurement in Radiating Environments 121


A.1 Additional Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . 121
A.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
A.3 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
A.4 Heat Transfer Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
A.5 Calibration Measurements . . . . . . . . . . . . . . . . . . . . . . . . 126
A.6 Furnace Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 128
A.7 CFD Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A.8 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 133

B Measurement Accuracy of Q̇solar 135

C Radiation Absorption in the Gas Phase 137

Bibliography 149
x Contents
Nomenclature

Acronyms

BC Boundary Condition
CCS CO2 Capture and Sequestration
CTL Coal to Liquids
CPC Compound Parabolic Concentrator
CCR Conradson Carbon Residue
ETH Eidgenössische technische Hochshule
IGCC Integrated Gasification Combined Cycle
FV Finite Volume
GCI Grid Convergence Index
LHV Lover Heating Value
PDF Probability Density Function
PSI Paul Scherrer Institut
RMS Root Mean Square
SEM Scanning Electron Microscope
VR Vacuum Residue

Latin Characters

A window absorptance
Ci concentration of species i
cp specific heat capacity, J/(kg · K)
D particle diameter, m
Dc diffusion coefficient, m2 /s
E emitted/absorbed power in the MC solver
EA apparent activation energy
xii Nomenclature

Eλb Planck’s blackbody spectral emissive power, W/(m2 · µm)


f particle size distribution function
fV solid volume fraction
h enthalpy, J/kg
h convective heat transfer coefficient, W/(m2 · K)
I radiation intensity, W/(m2 · sr)
i chemical species
K kinetic rate constant
k thermal conductivity, W/(m2 K)
k imaginary part of the refractive index
k0 kinetic frequency factor
ln liters under normal conditions: 273 K and 1 bar
M molar mass, kg/mol
m mass, kg
ṁi mass flow rate of species i, g/min
Np number of particles
n number of moles
n real part of the refractive index
ṅi molar flow rate of species i, mol/min
p pressure, Pa
Qa,s,ext absorption, scattering and extinction efficiencies for a single sphere
Q̇ heat rate, W
Q̇solar incident solar radiation, W
q̇ heat flux, W/m2
q̇solar incident solar radiative flux, W/m2
R universal gas constant, J/(mol · K)
Ri volumetric reaction rate of species i, kg/(m3 · s)
Rw window reflectance
r radial coordinate, m
ri mass specific reaction rate of species i, mol/(g · s)
S cross sectional area, m2
S surface area, m2
Si sensitivity
s distance along path
T temperature, K
Nomenclature xiii

Tr w window transmittance
t time, s
U overall heat transfer coefficient, W/(m2 · K)
u velocity, m/s
V volume, m3
V̇ volumetric flow rate, ln /min
XC carbon conversion
XH2 O steam conversion
x axial coordinate, m
Yi mass fraction of species i
yi molar fraction of species i

Greek Characters

β extinction coefficient, 1/m


χ fraction
∆HR reaction enthalpy, J/mol
 emissivity
 relative error
η particle effectiveness
ηchem solar-to-chemical energy conversion efficiency
ηproc solar thermal process efficiency
κ absorption coefficient, 1/m
µ dynamic viscosity, kg/(m · s)
ρ density, kg/m3
Φ scattering phase function
φ energy source term, W/(m3 · s)
σs scattering coefficient, 1/m
σ Stefan-Boltzmann constant 5.67051 · 10−8 , W/(m2 · K4 )
τ relaxation time, s
τ mean residence time, s
ξ size parameter
xiv Nomenclature

Subscripts/Superscripts

0 initial
a absorbed
b blackbody
chem used by the chemical reaction
cond conduction
conv convection
g gaseous phase
i species
in inlet
rad radiative
rerad reradiation from the cavity
s solid phase
sk shrunk
T total
w window
λ spectral
Chapter 1

Introduction

Today’s energy sector is facing the important challenge of meeting the rising demand
for energy without increasing the emission of greenhouse gases and, in particular,
those of CO2 . This increasing demand, combined with the predicted decline in
conventional oil production in 10-20 years [78], can be met by increased utilization
of coal, whose reserves are estimated to last for 250 years [82]. With the increase in oil
prices also the cost- and energy-intensive extraction of oil shales and tar sands [85], as
well as inter-fuel substitution (coal-to-liquids and coal-to-gas technology, CTL) [47,
58] have gained in importance. Unfortunately, these technologies do not contribute
to the reduction of CO2 emission. Clean coal technologies, including CO2 capture
and sequestration (CCS) [49] and increased plant efficiency, i.e., the use of coal-fired
combined cycle processes (IGCC and PFBC) [75], are proposed as countermeasures
to mitigate the ecological drawbacks of a coal-based energy economy [15].
A higher potential in meeting the mentioned challenges altogether is attributed
to the substitution of fossil fuels by clean renewable fuels on one side [46, 22] and
the enhancement of material and energy efficiency on the other [45]. Typical sub-
stite fuels are, among others, fuels produced from biomass (biofuels) [35, 51] and
fuels produced by means of solar energy (solar fuels), typically hydrogen [93]. The
complete substitution is, if ever achievable, a long term goal. Nonetheless, mid term
goals aimed at the development of hybrid solar/fossil technologies are of strategic
importance to the creation of a transition path toward solar fuels.
An important example of such a hybrid solar/fossil process, in which fossil fuels
are used exclusively as the chemical source for H2 production and concentrated solar
power is used exclusively as the energy source of process heat, is the endothermic
2 1. Introduction

steam-gasification of petroleum coke (petcoke) to synthesis gas (syngas). Petcoke is


a solid residue from the processing of heavy and extra heavy oils using delay-coking
and flexicoking technologies. The calorific value of the feedstock is upgraded by
solar power in an amount equal to the enthalpy change of the reaction. A Second-
Law analysis for generating electricity using solar gasification products indicates the
potential of doubling the specific electrical output and, consequently, halving the
specific CO2 emission, vis-à-vis conventional petcoke-fired power plants [112, 101].
The sun is an attractive source for primary energy. In fact, the solar energy re-
serve is virtually unlimited, freely available, and its utilization is ecologically benign.
On the other hand, the incident solar radiation on the earth surface is diluted (≈ 1
kW/m2 ), intermittent and unequally distributed. These intrinsic drawbacks of solar
energy, which affect in particular solar electricity generation, are overcome by solar
gasification and by solar thermochemistry, in general. Solar gasification is therefore
an attractive way to transform solar energy into a storable and transportable energy
carrier.

1.1 Thesis Outline


The present thesis is performed in the framework of a joint project between the
Swiss Federal Institute of Technology in Zurich (ETH) and the Paul Scherrer In-
stitut in Villigen (PSI), both Switzerland, the national research center ‘Centro de
Investigaciones Energéticas, Medioambientales y Tecnológicas’ in Spain (CIEMAT),
and the research and development center of Petróleos de Venezuela, S.A. (PDVSA /
INTEVEP). The main goal of the project is the development and demonstration of
a process and of the related technology required for the production of high quality
syngas from carbonaceous materials using a solar thermochemical process.
A previous PhD thesis was done by D. Trommer in the same framework [100].
His work focused on the thermodynamics of petroleum coke gasification, and on the
investigation of the chemical kinetic mechanisms involved. The work presented here
covers three additional tasks of the project: (1) the conception, design, manufactur-
ing and experimental operation of a lab-scale prototype reactor at a power level of 5
kW, (2) the numerical modeling of the process and of heat and mass transfer inside
the reactor, and (3) the design of a scaled-up version of the reactor to a power level
of 500 kW.
This thesis is divided into two main parts. Part I is dedicated to the experimen-
1.1. Thesis Outline 3

tal work (Chapters 2-4). The engineering design, the operation of such a reactor on
a solar furnace, the types and morphology of the employed feedstock and the ex-
perimental results are presented. Part II focuses on the modeling aspects (Chapters
5-8). Developed models were used to support the engineering design, to identify
parameters affecting the efficiency of the concept and to predict the performance of
the scaled-up reactor.
The characteristics and origin of the employed feedstock types are described
in Chapter 2. Elemental composition and particle size distributions are given for
Petrozuata Delayed Coke. Production and rheology of coke water slurry is shortly
addressed. Finally, the properties of vacuum residue are presented.
Chapter 3 focuses on the experimental setup used. The design of the cavity-
type reactor, the instrumentation and the outline of the concentrating facility are
described in detail. Furthermore, the key performance parameters of the process
are defined and their derivation and accuracy is discussed. These parameters are
the product gas composition, the energy efficiency, the chemical conversion and the
average reactant’s residence time in the cavity.
In Chapter 4 the results of three experimental campaigns are reported, each
one for a distinct feedstock: (1) Dry coke powder with steam fed separately (further
referred to as Campaign 1), (2) coke-water slurry injected into the cavity (Campaign
2) and (3) liquefied vacuum residue with steam fed separately (Campaign 3). Syngas
was successfully produced in all three campaigns. Differences in reactor performance
related to the feedstock type and injection system are addressed. Material from this
chapter has been published in [123, 122] and [121].
General consideration on the framework that supports the simulation models are
given in Chapter 5. The characteristics of the incoming radiation as well as the
conductive behavior of the reactor walls was determined by stand-alone preliminary
simulations. Wall and window radiative properties are presented. The concept of
‘equivalent monodisperse diameter’ is introduced to model the polydisperse feedstock
types, which have random size-distribution functions. Finally, the kinetic rate laws
for steam gasification of petcoke developed by Trommer [102] and employed in this
thesis are shortly presented.
In Chapter 6 a lumped parameters model — used for system analysis — is
described. It is used to support the early stages of the engineering design. The model
is validated with the experimental results. Good agreement is found especially for
the temperatures and carbon conversion rates of Campaign 1. Parameter studies
4 1. Introduction

are performed and the beneficial aspect of up-scaling are investigated.


In Chapter 7 a model for heat and mass transfer in the reactor’s cavity is pre-
sented. It solves simultaneously for convective, conductive and radiative heat trans-
fer as well as for mass transport and chemical kinetics. Focus is put on the radiative
heat transfer in participating media. The numerical implementation, including the
coupling of a simple finite-volume flow solver with a Monte-Carlo raytracer that
solves the equation of radiative transfer is described in detail. The model is suc-
cessfully validated with good agreement for the two campaigns considered (1 and
2 with petcoke as feedstock). Material from this chapter has been submitted for
publication [125].
Finally, in Chapter 8 the model developed and validated in Chapter 7 is used to
study in detail the parameters that affect the reactor’s performance. These are the
morphology of the feedstock, the feeding rates and the geometry of the cavity. The
same analysis is also performed for the scaled-up reactor. Better performance due
to lower heat losses as well as increased flexibility towards feedstock morphology are
predicted. Material from this chapter has been submitted for publication [126].
Appendix A is dedicated to an apparatus developed to measure gas temperatures
in highly radiating environments. It is based on a simple suction thermocouple
supplemented by a correction model that accounts for radiative, conductive and
convective heat transfer. This chapter has been published as [120].

1.2 Conventional Gasification


Gasification is a process that, under addition of a gasification agent, usually CO
or steam, converts any carbon-containing material, such as coal, petroleum, coke,
biomass or waste into a synthesis gas (syngas) composed primarily of carbon monox-
ide and hydrogen. This endothermic reaction takes place at temperatures above
approx. 1000 K. Since the 1850s gasification technologies have been applied com-
mercially. In between 1850 and 1940 gasification of coal was used to produce ‘town
gas’ for light and heat. Virtually all gas for fuel and light was produced by gasifi-
cation until the development of natural gas supplies and transmission lines in the
1940s and 1950s. During World War II German engineers used gasification to pro-
duce synthetic fuel from coal. This technology was exported to South Africa in
the 1950s, where it was further developed to produce liquid fuels and chemicals.
More recently, gasification has gained in importance as an environmentally-friendly
1.2. Conventional Gasification 5

disposal and conversion technology of residues from heavy crude oil processing. Con-
ventional gasification is also used in integrated gasification combined cycle (IGCC)
plants fired with low or negative-valued feedstocks for clean and efficient electricity
production [58, 79].
In conventional gasification the energy required to heat the reactants and to
run the chemical reaction is either produced by partial oxidation of the feedstock
(autothermal gasifiers) or provided externally (allothermal gasifiers). Autothermal
gasification has the advantage that any losses associated with the heat transfer
are avoided and the construction of the gasifier is simplified. Furthermore, the
temperature can be adjusted quickly and accurately by regulation of the amount
of oxygen injected. On the other hand, allothermic gasification usually results in
higher-quality syngas, because the products are not contaminated by combustion
byproducts. Gasification reactors are divided into three groups depending on how
the solid fuel is brought into contact with the gasification agent: fixed/moving bed,
fluidized bed or entrained flow. The type of reactor influences the residence time, re-
actor temperature and pressure, and certain characteristics of the produced gas. For
this reason each reactor type is suited for a specific type, rank, and size distribution
of solid feedstock.
Fixed/moving bed gasifiers contain a bed of lump fuel supported by a grate and
maintained at a constant height. The feedstock is fed from the top end and flows
countercurrent to the rising gas stream. A single particle moving through the bed
passes different zones including drying and preheating, devolatilization, gasification,
oxidation, and ash removal. Fixed bed gasification systems are simple, reliable and
offer high efficiency with respect to feedstock and energy consumption. Outlet gas
temperature are in the range 425–650 ◦ C, feedstock size is between 6 and 50 mm. As
a consequence of the moderate temperatures the product gas contains around 10%
hydrocarbons, mostly CH4 . A typical example of a commercial fixed bed gasifier is
the dry ash Lurgi gasifier [76, 37]. Lurgi gasifiers were employed on a large scale for
South Africa’s SASOL complex [21].
Fluidized bed gasifiers accept feedstock as grains with a size in the range 6 to
10 mm. The solid feedstock is suspended on upward-blowing jets of the gasification
agent. The result is turbulent mixing of gas and solids. The tumbling action, much
like a bubbling fluid, provides more effective chemical reactions and heat transfer.
Reactors of this type are characterized by outlet gas temperatures in the range 900 to
1050 ◦ C, a high specific gasification rate and product uniformity. A typical example
6 1. Introduction

of the fluidized bed technology is the Winkler process [10, 37].


Entrained-flow gasifiers are operated with pulverized feedstock with particles of
less than 100 µm. The solid feedstock is entrained with the gasifying agent to react
in a concurrent flow having the form of a high temperature flame. The principal
advantages of this process are the ability to handle practically any coal as feedstock
and to produce a clean, tar-free and low-in-methane synthesis gas. Entrained-flow
reactors require relatively high flow rates leading to small residence times. In order
to obtain full conversion of the coke, entrained flow reactors operate at very high
temperatures (above 1400 ◦ C) and require, therefore, a comparably high oxidant
supply. Devolatilization products are released in the high temperature region and
thus further cracked and oxidized. This reactor type is usually operated at 20-70
bar [37, 58]. This technology has bee selected for the majority of commercial IGCC
plants. The two best-known types of entrained-flow gasifiers are the top-fired coal-
water-slurry feed gasifier, as used in the Texaco process [87] and the dry coal feed
side-fired gasifier as developed by Shell and Krupp-Koppers (Prenflo)[109, 37].

1.3 Solar Thermochemistry and Solar Gasifica-


tion in Particular
Solar thermochemistry refers to a number of process technologies that harness con-
centrated solar energy by absorbing sunlight in an endothermal chemical reaction
occurring at high temperatures. The products of these chemical reactions are usually
called solar fuels or solar energy carriers, as they carry solar energy in an amount
equal to the reaction’s enthalpy change. Overviews on the different technologies en-
visaged and on processes that were experimentally demonstrated are given in [52, 28]
and [94]. These processes can be divided into three groups based on the employed
feedstock and on the chemical reactions involved: (1) direct thermal splitting of
water or H2 S, a byproduct from natural gas, petroleum or coal refining, (2) H2 O
splitting thermochemical cycles and (3) upgrading and decarbonization of fossil fuels.
Despite being a simple and intuitive way to produce hydrogen, the direct thermal
splitting of water is difficult to achieve in practice. In fact, the temperatures required
for an efficient dissociation lie above 2500 K. Furthermore, the explosive mixture
of gaseous products (H2 and O2 ) needs to be separated at high temperatures to
avoid recombination. The feasibility of the process has been investigated among
1.3. Solar Thermochemistry and Solar Gasification in Particular 7

others by [27] and [54]. An alternative way to produce hydrogen by direct thermal
splitting is the decomposition of H2 S, a highly toxic industrial byproduct recovered
in the sweetening of natural gas and in the removal of organically-bound sulfur
from petroleum and coal. The process runs at temperatures around 1800 K on a
Al2 O3 surface. Separation of the products H2 and S2 occurs by condensation and
subsequent solidification of the sulfur [48].
In order to reduce the reaction temperature and eliminate the need for high-
temperature gas separation, various processes for the production of hydrogen by
multi-step thermochemical water splitting cycles have been proposed — typically
two-step processes based on metallic redox systems. In the first step the metal oxide
is partly or completely reduced in a solar high-temperature endothermic reactor; in
the second step the product from the first step is oxidized with water in an exother-
mal reaction releasing hydrogen (hydrolysis reaction) closing the cycle. An overview
of the most considered redox pairs is given in the following. The Fe3 O4 /FeO cycle
was recently experimentally investigated by [16]. Complete thermal dissociation of
Fe3 O4 at 1973 K and 0.1 bar under inert atmosphere was reported, while up to
80% conversion at 850 K was obtained for the hydrolysis step. Systems based on
Mn3 O4 /MnO, Co3 O4 /CoO can be thermally decomposed in air at 1810 and 1175 K,
respectively. Unfortunately, the H2 yields obtained were only 0.002% and 4 · 10−7 %
at 900 K for Mn3 O4 /MnO and Co3 O4 /CoO, respectively. High H2 yield of 99.7% at
900 K was obtained for Nb2 O5 /NbO2 , but the thermal decomposition temperature
of 3600 K in air is prohibitively high [62]. Extensive research has been performed on
the ZnO/Zn cycle. Latest results for direct thermal dissociation of ZnO are reported
by [89]. The in-situ formation and hydrolysis of Zn nanoparticles was investigated
by [23]. The issues arising from the high temperatures required for the direct ther-
mal dissociation of ZnO (approx. 2000 K at 1 bar) are mitigated by the addition
of carbon to the feedstock at the expense of CO2 emissions. Solar carbo-thermal
reduction of ZnO was successfully demonstrated at a power level of 10 kW [70] and
at 300 kW [115]. The produced zinc was reported to be significantly more reac-
tive in the hydrolysis step than commercially available material [107]. Finally, a
three-step sulfur-iodine cycle is proposed by [81], in which sulphuric acid is split by
concentrated solar radiation.
The processes involved in upgrading and decarbonization of fossil fuels can be
grouped into three categories: (1) solar thermal cracking of hydrocarbons, in which
solid carbon and hydrogen is produced, (2) solar reforming, in which a gaseous
8 1. Introduction

hydrocarbon is reacted to a H2 -CO mixture (syngas) in a steam or CO2 atmosphere,


and (3) solar steam- or CO2 -gasification of solid carbonaceous materials [111]. High
temperature solar pyrolysis of coal and biomass was already reported in 1983 by [7]
and [4], respectively. The solar pyrolysis of oil shale and the recovery of the volatile
product was reported by [8]. Lately, research focused on thermal splitting of methane
into hydrogen and carbon black. Successful lab-scale experiments were reported by
[41], [18] and [1]. The solar reforming of methane was reported among others by [19]
without catalyst, by [53] on a catalytic Ru/Al2 O3 surface, and by [13] on a catalytic
rhodium surface.
Successful solar gasification of carbonaceous materials was first reported in the
80’s: coal, activated carbon, coke, and coal/biomass mixtures were employed in a
fixed bed windowed reactor by [31]. Charcoal, wood and paper was gasified with
steam in a fixed bed reactor, and charcoal was gasified with CO2 in a fluidized
bed by [95]. More recently, [26] reported on gasification of oil shale and coal in a
fixed bed reactor, [65] on gasification of waste tyres and plastic (PET), and [112]
on gasification of coal in a fluidized bed reactor. Lately, extensive studies on the
steam gasification of petroleum coke and vacuum residue were performed. Chemical
kinetics were measured by [101] and three distinct feedstocks were processed in a
lab-scale entrained-flow reactor: dry coke particles [123], coke-water slurries [122]
and vacuum residue [121].
Part I

Experimental Work
Chapter 2

Feedstock

The materials used as feedstock for the solar gasification experiments presented in
this thesis are two typical residuals of heavy crude oil refining: a vacuum residue
(VR) and a delayed petroleum coke (coke). These are produced from the processing
of extra heavy crude oil from the Petrozuata oil field in the Orinoco belt (southern
strip of the eastern Orinoco river basin in Venezuela) and were provided by PDVSA,
the Venezuelan oil company.

crude oil distillation(fractionation) delayed coking


coke
gas
atmosperic
distillation

naphtha
gasoline gas/gasoline
desalter
kerosene
furnace diesel
coke drum

coke drum

heavy gas oil


gas oil
fractinator

atmospheric
residue
distillation

vacuum gas oil


vacuum

lubricating oil
furnace
furnace

vacuum
residue (VR) residue recycle

Figure 2.1: Schematic representation of a refinery. Only units involved in the con-
version of crude oil to vacuum residue and delayed coke are shown.
12 2. Feedstock

Figure 2.1 shows the main components of a refining facility that are involved in
the production of either VR or coke. Further processing of volatile hydrocarbons is
omitted, whereas the focus is put on the treatment of the residues of the distillation
and coking processes.
Crude oil is usually contaminated with water, inorganic salts, suspended solids,
and water-soluble trace metals. In order to reduce corrosion, plugging, and fouling
of equipment, and to prevent poisoning of the catalysts it is, therefore, desalted
(cleaned) prior to further processing. Crude oil refining is a complex process, which
involves numerous specific secondary processes. Nevertheless, it can be split into
two major steps, especially with respect to the production of vacuum residue and
coke.

Distillation Desalted crude oil is separated into various fractions by distillation


(fractionation), first in an atmospheric column and subsequently in a vacuum column
operated at 30–50 mbar. The temperatures required for distillation of the heavier
fraction of crude oil at ambient pressure are so high that cracking would occur.
These materials are, therefore, distilled under vacuum, since a lowering of the pres-
sure results in lowering of their respective boiling temperatures. Partial pressures
are further lowered by steam injection. The main fractions obtained have specific
boiling-point ranges and can be classified in order of decreasing volatility into gases,
light distillates, middle distillates, gas oils, and vacuum residuum (VR). In contrast
to the distillation step, where thermal cracking is avoided, further processing of the
VR aims at the destruction of the long hydrocarbon chains of the heavy molecules.

Coking Heavy residuals are upgraded into lighter products or distillates by a


severe method of thermal cracking (coking). Coking produces straight-run gasoline
(coker naphtha) and various middle-distillate fractions used as catalytic cracking
feedstock. The final residue is a form of carbon called coke. The two most common
processes are delayed coking and continuous (contact or fluid) coking.
In delayed coking the charge, typically atmospheric or vacuum residues, is heated
above the coking point (480–520 ◦ C) and subsequently fed to a large coke drum,
which provides the long residence time (24 hours) needed to allow the cracking
reactions to proceed to completion. Premature coking in the heater tubes can be
avoided by sufficiently high flow rates. Vapors from the drums are returned to
a fractionator where gas, naphtha, and gas oils are separated out. The heavier
2.1. Petrozuata Delayed Coke 13

hydrocarbons produced in the fractionator are recycled through the furnace. In


order to maintain continuous operation usually two coke drums are used, one on
stream and the other being emptied. The full drum is steamed to strip out uncracked
hydrocarbons, cooled by water injection, and decoked by mechanical or hydraulic
methods.

2.1 Petrozuata Delayed Coke

Table 2.1: Coke feedstock types used in the experiments.

type shipment Dmin –Dmax 1 , µm D30 2 , µm


1 1 jet milled 0.51–17.4 2.21
2 1 ball milled 0.76–300 6.7
3 2 sieved, 80 µm screen 1.5–300 17.6
4 2 sieved, 200 µm screen 2.3–592 30.8
5 2 as received 3–678 40.0
1
Values strongly affected by the detection limit of the Horiba LA950.
2
See eq. (5.10) for the definition of the mean diameter D30 .

The raw coke provided by PDVSA was further processed to the five distinct types
of feedstock listed in Table 2.1. A first shipment consisted of relatively big particles
with characteristic diameters of a few centimeters. Feedstock types 1 and 2 were
obtained by grinding the raw material with a jet mill and a ball mill, respectively
(ARP GmbH, Loeben, Austria). A second shipment was composed by somewhat
smaller particles with a maximal diameter around 1 mm. This material was sieved
with a 80 µm and a 200 µm screen to produce feedstock types 3 and 4, respectively.
Type 5 describes the raw material as received.
Figure 2.2 shows the particle size distribution functions measured by laser scat-
tering with a Horiba LA950. Plotted are the number density f (D) and the respective
volume density f (D)·D3 for the five types of feedstock listed in Table 2.1. Although
the relative number of big particles seems negligible from Fig. 2.2 (a), those parti-
cles carry an important share of the total mass as shown by the volume distribution
curves shifted to the right in Fig. 2.2 (b). SEM pictures of feedstock type 1 are
shown in Fig. 2.3. Frames (a) and (b) show the unreacted spherically-shaped par-
ticles with diameters in the micrometer range. Frames (c) and (d) show particles
after pyrolysis above 1400 K, in which release of volatile matter and thermal cracking
14 2. Feedstock

(a) (b)

1
2 1

volume density f (D) · D3


number density f (D)

5 2 3 4
5
3 4

1 2 5 10 20 1 2 10 100 500
particle diameter D, µm particle diameter D, µm

Figure 2.2: Particle size distribution functions of the petcoke feedstock types listed
in Table 2.1. (a) shows the number density (also called population density); (b)
shows the volume density. Type 1: jet milled, type 2: ball milled, type 3: sieved
with a 80 µm screen, type 4: sieved with a 200 µm screen, type 5: as provided by
PDVSA.

result in the formation of small structures on the particles surface. Finally, frames
(e) and (f) show particles after a typical experimental run from the first campaign
with combined pyrolysis and gasification and carbon conversion of XC =0.75. As
expected, the size of these particles is smaller compared to Fig. 2.3 (a) and (b).
Furthermore, and in contrast to the purely pyrolyzed feedstock, no formation of
small structures on the surface is observed. The elemental composition of the PD
coke is shown in Table 2.2. It mainly consists of carbon (62% molar) and hydrogen
(35% molar) with minor shares of nitrogen, oxygen and sulfur together with nickel,
vanadium and sodium in the ppm range. The lower heating value (LHV) was calcu-
lated as the sum of the elemental LHVs, since the exact composition of the bound
hydrocarbons is not known. The obtained value of 35.8 MJ/Kg is slightly higher
than the LHVs reported for coal [79] (28.5–35.3 MJ/Kg).
2.1. Petrozuata Delayed Coke 15

(a) (b)

1¹m 1¹m

(c) (d)

1¹m 1¹m

(e) (f)

1¹m 1¹m

Figure 2.3: SEM micrographs of petcoke samples used in Campaign 1 (feedstock type
1). (a) and (b) show unreacted particles, (c) and (d) show particles after pyrolysis
above 1400 K, (e) and (f) show particles after a typical experimental run with
combined pyrolysis and gasification and carbon conversion XC =0.75. Magnification:
(a),(c) and (e) 4’000 x; (b),(d) and (f) 10’000 x
16 2. Feedstock

Table 2.2: Approximate main elemental chemical composition (ultimate analysis),


low heating value and molar ratios H/C and O/C for PD coke.

wt% mol%
carbon 88.21 62.06
hydrogen 4.14 34.71
nitrogen 2.28 1.37
oxygen 1.46 0.77
sulfur 4.16 1.10
nickel ppm 414
vanadium ppm 2207
sodium ppm 100
LHV kJ/kg 35876
H/C mol/mol 0.5581
O/C mol/mol 0.0124

2.1.1 Characterization of Coke-Water Slurries3


In the second experimental campaign presented in Section 4.2 coke particles-water
mixtures, so called slurries, were employed. Coal water slurries (CWS) are well
known as an alternative to petroleum fuels used in furnaces and kilns for building
material industry and in industrial and heating boilers [90]. Typical coal-water
slurries consist of 60–75% coal and 25–40% water, with particle sizes in the range
of 50–500 µm. Low fractions of water are desirable in order to keep the reduction
in the fuel’s LHV small. Low viscosity is ensured by the addition of typically 1% of
chemical additives [20]. Sedimentation stability, rheology and the effect of particle
size on the slurry properties were investigated among others by [105] and [97].
The negative effect of excess water is less important for the steam gasification
process analyzed in this thesis. In fact a minimum of 57% wt water is required to
obtain a stoichiometric H2 O/C mixture of the reactants (additional water beyond
57% wt in turn reduces the efficiency). Two types of slurry where analyzed: sample
1 was made of coke particles with an average diameter of 2.21 µm compared to
17.6 µm for sample 2 (see Table 2.1). The water-to-carbon molar ratio was 2/1
and 1/1, corresponding to mass ratios of 73%/27% and 57%/43% for sample 1 and

3
Material presented in this section has been developed in the framework of: F. Kritter. Injection
system of coal water slurries for solar reactors. Diploma thesis, ETH Zurich, 2005.
2.1. Petrozuata Delayed Coke 17

(a) (b)
700 140
sample 1
dynamic viscosity µ, m · Pa/s

600 sample 2 120

shear stress τ , Pa
500 100

400 80

300 60

200 40

100 20

0 0
0 500 1000 1500 0 500 1000 1500
shear rate γ, 1/s shear rate γ, 1/s

Figure 2.4: Viscosity (a) and shear stress (b) as a function of the shear rate of a
water-coke slurry, measured for two types of feedstock [57]. The arrows indicate
the upward and downward curves, obtained by increasing and subsequently decreas-
ing the shear rate, respectively. Nicely visible is the hysteresis loop, typical for
thixotropic fluids.

sample 2, respectively. The slurries were prepared without any chemical additives
by mechanical mixing for several minutes. Viscosity and shear stress were measured
by shear rheometry at CIEMAT in Madrid. Results are plotted in Fig. 2.4. Curves
for dynamic viscosity and shear stress as a function of the shear rate, first increased
from 0 to 1500 1/s and subsequently decreased to 0 1/s, are show in Fig. 2.4 (a)
and (b) respectively. The shape of the curves indicates non-Newtonian thixotropic
behavior of the slurries. This means that the slurries undergo a decrease in viscosity
for increased shear rates. For example, for a shear rate of 250 1/s the viscosity
was 270 and 98.5 m · Pa/s, whereas at a shear rate of 1000 1/s it was 106 and 40.5
m · Pa/s for sample 1 and sample 2, respectively. As expected, the smaller particles
form a more viscous slurry, even if the water-to-coke ratio is doubled. Uniform
slurries were easily produced with the bigger particles but a the cost of storage
instability. In fact, separation of the two phases was observed after a short period
of time, especially for sample 2. During the experimental runs in the solar furnace
the slurry tank was, therefore, constantly stirred.
18 2. Feedstock

2.2 Petrozuata Vacuum Residue


The vacuum residue delivered by PDVSA is a residue from processing of extra-heavy
crude oil from the Petrozuata fields in Venezuela. It is solid at ambient temperature,
has plasto-elastic behavior for moderate shear rates but breaks into little pieces at
high shear rates, for example when hit with a hammer, showing a brittle fracture
surface. The VR liquifies at around 120 ◦ C, and strong release of volatile matter is
observed above 200 ◦ C (visual observations). Coking, that is the transformation to
a solid by thermal cracking, is reported to take place approx. above 400 ◦ C and has
to be avoided by any means in the feeding system. The elemental composition of
vacuum residue is given in Table 2.3 (data from PDVSA/INTEVEP). Significantly
larger amounts of hydrogen are detected compared to PD coke (57% vs. 37 % molar),
thus resulting in an increased LHV (as for coke, the LHV was calculated as the sum
of the elemental LHVs). On the other hand, smaller amounts of nitrogen, oxygen,
sulfur and metals are detected.
Table 2.3: Approximate main elemental chemical composition (ultimate analysis),
low heating value and molar ratios H/C and O/C for vacuum residue.

wt% mol%
carbon 86.19 42.52
hydrogen 9.6 56.84
nitrogen 0.85 0.36
oxygen 0.16 0.059
sulfur 3.1 0.57
nickel ppm 185
vanadium ppm 774
LHV kJ/kg 40908
H/C mol/mol 1.337
O/C mol/mol 0.0014
Chapter 3

Experimental Setup

An experimental reactor was developed to demonstrate the feasibility of solar steam


gasification of extra heavy crude oil or derived residues such as vacuum residue or
coke. The reactor was successfully tested at PSI’s solar furnace for 3 distinct exper-
imental campaigns underlining the flexibility of the implemented design and serving
as valuable support to the process of up-scaling. Feedstock was injected as coke
powder (Campaign 1), coke-water slurry (Campaign 2) or liquefied vacuum residue
(Campaign 3). This chapter describes the general experimental setup, including the
reactor design, details about the injection systems and a description of the facility
were the experiments were performed.

3.1 Reactor Design


The engineering design of the prototype reactor was based on a set of lab-scale
preliminary studies. Semi-batch mode experiments (a batch of feedstock flushed
with a continues gas stream) were performed in a conventional electrical furnace [2]
and [92], with a simple quartz dome reactor [25] and with an opaque ceramic tube
[3] mounted in the focus of ETH’s solar simulator [42]. Evaluated reactor concepts
included also direct/indirect concentrated solar irradiation, packed/fluidized beds
and semibatch/continous feeding. The highest potential was identified for a cavity
type reactor, where the constantly fed particulate feedstock is directly irradiated
while it is entrained by the reactive flow.
The final reactor configuration — developed in a joint effort at ETH in coop-
eration with engineers from the project partners — is schematically shown in Fig.
20 3. Experimental Setup

reactor insulation

tc S

product outlet
(Al2O3 / ZrO2)
reactor shell

reactor liner
(Inconel)

(Al2O3)
tc K
water cooled window fitting

tc K
multifunctional inlet
oil cooled frustrum

tc K

tc K
tc K

tangential injection
ring gap

tc S
tc K

nozzles

tc K
purging
nozzles

aperture

window

Figure 3.1: Scheme of the solar chemical reactor configuration used for the steam
gasification of carbonaceous materials at a power level of 5kW. The location of the
thermocouples (tc) of types K and S are indicated by stars.

3.1. It consists of a cylindrical cavity-receiver, 210 mm in length, 120 mm inside


diameter, that contains a 5 cm-diameter opening — the aperture — to let in con-
centrated solar power. The cavity-type geometry is designed to effectively capture
3.1. Reactor Design 21

the incident solar radiation; its apparent absorptance is estimated to exceed 0.95
[68]. The cavity is made out of Inconel 601, lined with Al2 O3 , and insulated with
an Al2 O3 /ZrO2 ceramic foam. The aperture is closed by a 0.3 cm thick, clear fused
quartz window, mounted in a water-cooled aluminum ring that also serves as a shield
for spilled radiation. The window is actively cooled and kept clear from particles
and/or condensable gases by means of an aerodynamic protection curtain created
by tangential flow through four tangential nozzles combined to radial flow through a
circular gap. In front of the aperture, the cavity-receiver is equipped with a diverg-
ing conical funnel for mounting the window 6 cm in front of the focal plane, where
the radiation intensity is about 10 times smaller and dust deposition is unlikely to
occur. Since radiation spillage can reach flux concentration ratios greater than 1000,
corresponding to an incident flux of 1000 suns (1 sun = 1 kW/m2 ), this component is
actively oil-cooled and kept in the range 393–453 K to prevent steam condensation.
Steam and particles are injected separately into the reactor cavity, permitting
separate control of mass flow rates and stoichiometry. Steam is introduced through
several ports. Based on flow visualization experiments in Plexiglas models and CFD
simulations, best flow patterns in terms of residence time and flow stability were
obtained with two sets of 4 symmetrically distributed tangential nozzles, located in
planes 4 and 12 cm behind the aperture plane, as shown in Figure 3.1. The petcoke
feeding unit is positioned on the top of the reactor vessel with its inlet port located
at the same plane as the primary steam injection system, allowing for the immediate
entrainment of particles by the steam flow. Inside the cavity, the gas-particle stream
forms a vortex flow that progresses toward the rear along a helical path. Reactor
wall temperatures were measured in 12 locations with type K thermocouples, in-
serted in the Inconel walls and not exposed to direct irradiation (indicated by the
stars in Fig. 3.1). The temperature of the inner Al2 O3 cavity was measured with a
solar-blind pyrometer that is not affected by the reflected solar irradiation because
it measures in a narrow wavelength interval around 1.39 µm where solar irradia-
tion is mostly absorbed by the atmosphere [103]. The nominal reactor temperature,
denoted Tcavity , was then calculated as the mean between the pyrometer reading (cor-
rected for window transmittance, Tpyrometer = Tpyrometer,reading · Tr −0.25
w , where Tr w is
the transmittance) and the thermocouple reading (corrected for conduction through
the Al2 O3 liner with eq. (6.9)). The exit gas temperature was measured with a
type S thermocouple in the outlet pipe, while the temperature of the inlet port was
monitored with an additional type K thermocouple.
22 3. Experimental Setup

3.2 Feedstock Injection


The reactor was primarily designed for processing of dry petcoke particles in the
10–100 µm range, in which particles and steam are injected separately. The inves-
tigation of alternative feedstocks, such as water-coke slurries and liquefied vacuum
residue, led to substantial changes in the design of the injection systems in order
to accommodate for their specific properties, while still creating a particle-laden
entrained flow.

3.2.1 Dry Powder Feeding

petcoke particles/ steam


argon tangential injection
nozzles

steam reactor insulation


(Al2O3 / ZrO2)
reactor shell
(Inconel)
reactor liner
(Al2O3)

steam

steam

Figure 3.2: Feedstock injection setup used in Campaign 1. Petcoke particles are fed
by means of a brush conveyer and further entrained by the steam injected from four
tangential nozzles.

Figure 3.2 shows the injection system used in the first experimental campaign.
The petcoke particles were stored in a tube and conveyed toward a rotating brush
by a piston. Argon was used to flush this rotating brush to ensure constant feeding
rates. The feeding unit was mounted on top of the reactor permitting the particles
to fall into the cavity. They were then entrained by the steam injected from four
tangential nozzles mounted at the same axial position. Smooth overall operation was
3.2. Feedstock Injection 23

experienced with this setup, and a homogenously distributed particle suspension was
obtained (visual observations), which lead to good reactor performance.

3.2.2 Slurry Feeding

petcoke/water slurry
particles/steam
injection (long)
reactor insulation
(Al2O3 / ZrO2) particles/steam
injection (short)
reactor shell
(Inconel)
reactor liner
(Al2O3)

Figure 3.3: Feedstock injection setup used in Campaign 2. The coke-water slurry is
preheated in the tubes cast into the reactor’s insulation, the water evaporates and
the particles are injected into the cavity.

Figure 3.3 shows the injection system used in the second experimental campaign.
The petcoke particles were mixed with water in a constantly stirred tank mounted
above the reactor. The coke-water slurry was then conveyed with a peristaltic pump
and introduced into the cavity trough an Inconel tube cast into the reactor liner. The
water in the slurry is thereby evaporated, the steam expands and the preheated feed-
stock is injected into to cavity. The longer original design of the injection tube had
to be substituted for a shorter one. In fact, due to the high temperatures achieved,
clogging by coking of the particles to the tube walls occurred. As a consequence of
the thixotropic behavior of the slurry the tubing had to be cleaned and filled with
water prior to each feeding cycle. In addition, clogging was prevented by the use of
bigger particles and excess water, which resulted in a lower slurry viscosity.
24 3. Experimental Setup

3.2.3 Liquefied Vacuum Residue

liqui¯ed VR steam/argon
VR droplets

reactor insulation release of


(Al2O3 / ZrO2) volatiles and
particle
reactor shell formation
(Inconel)
reactor liner deposition of
big particles
(Al2O3)

Figure 3.4: Feedstock injection setup used in Campaign 3. The liquefied vacuum
residue is injected through a cold nozzle, the droplets are subsequently heated, py-
rolyzed and gasified.

Figure 3.4 shows the injection system used in the third experimental campaign.
The vacuum residue is liquefied in a heated tank mounted above the reactor and
kept at approx. 423 K. The VR is conveyed by means of a gear pump, while both,
the tubing and the pump head are heated to approximately 393 K. Five different
reactant injection arrangements were tested, including a setup similar to the one used
in the slurry campaign. Clogging by coking was observed for temperatures above
673 K. The injection location was therefore moved out of the directly irradiated
region and further cooled by a coaxial steam/argon flow. In addition, the nozzle
was removed to ease the VR injection because of the insufficient pressure delivered
by the gear pump. After injection the VR droplets fall inside the cavity, are exposed
to the concentrated radiation and thereby release the volatile matter and break up to
form particles. Due to this low-level injection technology only a part of the resulting
particles was entrained by the gas flow, whereas the bigger particles deposited in the
cavity, thus lowering the reactor’s performance.
3.3. Peripherals and Facility 25

3.3 Peripherals and Facility

heliostat shutter reactor parabolic concentrator

Figure 3.5: Photograph of PSI’s high flux solar furnace. Main visible features are the
sun tracking heliostat, the Venetian-blind type shutter used to control the incoming
radiation and, inside the building, the parabolic concentrator and the reactor.

Experimentation was carried out at the PSI’s solar furnace [34]. This solar re-
search facility, shown in Fig. 3.5, consists of a 120 m2 sun-tracking heliostat in-axis
with an 8.5 m-diameter paraboloidal concentrator, and delivers up to 40 kW at peak
concentration ratios exceeding 5000 suns. A Venetian blind-type shutter located be-
tween the heliostat and the concentrator controls the power input to the reactor.
Radiative solar flux intensities were measured optically with a calibrated CCD cam-
era by recording the image of the sun on a water-cooled Al2 O3 -coated Lambertian
(diffusely reflecting) plate positioned at the focal plane. The reactor itself was sub-
sequently positioned with its aperture at the focal plane and intercepting the regions
of maximum solar flux intensity. Integration of the incident radiative flux over the
reactor’s aperture yielded the solar power input, Q̇solar . The accuracy of the opti-
cal measurement combined with the reactor’s misalignment led to an error of ±913 %
for Q̇solar (for details see Appendix B). Figure 3.6 shows an overview of the most
important peripherals attached to the reactor. Inlet gas flows of Ar and H2 O were
controlled using electronic flow meters (Bronkhorst HI-TEC). Solid particles were
26 3. Experimental Setup

Campaign 1 (2004) Campaign 2 (2005) Campaign 3 (2006)


coke H2O coke liquefied
particles particles VR residue

brush mixer
feeder
slurry
screw peristaltic gear pump
conveyor pump

oil cooling circuit optional


feedstock/steam solid
injection recirculation

solid residue
pyrometer

flare
concentrated
solar radiation cyclone

product
outlet
window purge

filter /
water cooling circuit condenser

Siemens
calomat
evaporator
Ar Siemens
H2O steam line IR IR
H2O detector
Ar Varian
Bronkhorst steam generator micro GC GC

Figure 3.6: Main components of the experimental setup at PSI’s high flux solar fur-
nace. Three different feeding systems are shown for the three distinct experimental
campaigns performed from 2004 to 2006.

removed from the outlet stream in a cyclone and excess moisture was condensed in a
counterflow heat exchanger. Product gases were analyzed on-line by gas chromatog-
raphy (Varian Micro GC, equipped with a Molsieve-5 and a Poraplot-U column; 1
ppm detection limit; 0.75 min−1 sampling rate). Better temporal resolution was ob-
tained by additional IR-based detectors for CO, CO2 , and CH4 (Siemens Ultramat
23; 0.2% detection limit; 1 s−1 sampling rate), and a thermal conductivity-based
3.4. Key Process Parameters 27

detector for H2 (Siemens Calomat 6; 50 ppm detection limit; 1 s−1 sampling rate
and interfering gas correction). Finally, the product gas was burned on a torch
mounted outside of the building. The reactor’s pressure was monitored with pres-
sure transducers while a pressure safety valve prevented overpressure derived from
a five-fold volumetric growth due to gas formation and thermal expansion. Three
distinct feeding systems were employed for each specific feedstock employed. The
feeding rate of dry coke particles was determined by measuring the weight differ-
ence of the particle conveyor. The rates for slurry and liquefied vacuum residue
were determined by calibration of the feeding pumps before the experiments. An
accuracy of ±5% was estimated. Steam was generated externally in a Bronkhorst
steam generator that allows H2 O concentrations in Argon up to 100%. A combined
heater/cooler was used to temperate the oil flowing trough the diverging frustum.
The spilled radiation impinging on the window mounting was removed by water
cooling. Representative solid product samples collected at the filter downstream of
the reactor were examined by scanning electron micrography.

3.4 Key Process Parameters

3.4.1 Product Gas Composition


The measuring equipment used to determine the gas composition at the outlet pro-
vides only relative values, namely the molar fractions yi of species i. The molar
production rate of a specific species ṅi is then given by:

ṅi = ṅtot · yi (3.1)

The total molar flow rate ṅtot at the exit of the reactor was calculated by means of
the known argon input flow rate ṅAr :

ṅAr
ṅtot = (3.2)
yAr

where yAr is the molar fraction of argon in the product gas estimated with yAr =
P
1 − yi , with i = H2 , CO, CO2 , CH4 and O2 under the assumption that argon is
the only non-negligible gas species not detected by the equipment.
28 3. Experimental Setup

A more accurate technique was introduced in Campaigns 2 and 3 to overcome


potential inaccuracies. Nitrogen was injected at a defined rate ṅN2 into the product
gas stream after the reactor. The total molar flow rate reaching the measurement
equipment is then ṅ0tot = ṅtot + ṅN2 , and can be calculated as ṅ0tot = ṅN2 /yN2 . The
total molar flow rate at the exit of the reactor ṅtot , required in eq. (3.1), can then
be expressed as:  
ṅN2 1
ṅtot = − ṅN2 = ṅN2 · −1 (3.3)
yN2 yN2
where yN2 is directly measured with the Varian Micro GC. For the purpose of cross-
checking a modified version of eq. (3.2) is used in Campaigns 2 and 3, which accounts
for the additional nitrogen flow:

ṅAr
ṅtot = − ṅH2 (3.4)
yAr

For example, comparison of the values for Campaign 2 revealed an average difference
of 6% with peak relative errors up to 22%. Note that the automatic interfering gas
correction of the Calomat does not work for nitrogen. The reading for the hydrogen
was therefore corrected manually by 0.09 % H2 for each % N2 .
Other gas species, such as H2 S, C2 H2 , C2 H4 and C2 H6 were not considered here
since their concentrations are negligible for the calculations of energy efficiencies and
chemical conversions.

3.4.2 Chemical Conversion


The share of carbon in the feedstock consumed by the gasification reaction, further
referred to as carbon conversion, was calculated by means of a C balance:

ṅCO + ṅCO2 + ṅCH4


XC = (3.5)
ṅ0C

where ṅ0C denotes the initial amount of carbon fed into the reactor. Similarly, water
conversion is obtained from the oxygen balance:

ṅCO + 2 · ṅCO2
XH2 O = (3.6)
ṅ0H2 O

where ṅ0H2 O denotes the molar amount of water fed by the steam generator and
water contained in the coke-water slurry fed in Campaign 2. Note that the excess
3.4. Key Process Parameters 29

steam, condensed right at the reactor outlet, can then be estimated as ṅH2 O =
(1 − XH2 O ) · ṅ0H2 O .

3.4.3 Energy Efficiency


The reactor’s thermal performance is described by two energy conversion efficiencies.
The solar-to-chemical energy conversion efficiency ηchem is defined as the portion of
solar energy input stored as chemical energy:

XC · ṅ0C · ∆HR |298K


ηchem = (3.7)
Q̇solar

where XC · ṅ0C describes the amount of carbon gasified and ∆HR |298K = 131 kJ/molC
is the reaction enthalpy. The solar thermal process efficiency ηproc is defined as the
portion of solar energy stored both as LHV in the product gas and sensible heat
— which potentially can be recovered — and takes into account the heat required
for steam generation (evaporation and superheating to 423 K) and the LHV of the
feedstock:
Pspecies R Treactor
ṅH2 LHV H2 + ṅCO LHV CO + i 473k
ṅi cp,i (T ) dT
ηproc = (3.8)
Q̇solar + Q̇steam + ṁfeedstock LHV feedstock

Theoretical maximal values for ηchem are calculated assuming perfect insulation
(Q̇cond = 0), complete chemical conversion (XC = 1) and stoichiometric steam-
to-carbon feeding rates. The useful energy available to heat and react the feed-
stock is then given by the solar power input diminished by the reradiation losses:
4
Q̇solar · (1 − σT
CI
), where C is the average concentration at the aperture and I is the
incident solar radiation intensity. The maximal molar feeding rate of carbon, ṅC ,
heated to a given reactor temperature T and completely reacted results then in

σT 4
Q̇solar · (1 − CI
)
ṅC = RT (3.9)
∆HR |298K + T∞
(cp,C + cp,H2 O + χ · cp,H2 ) dT

where χ is the hydrogen-to-carbon ratio 0.5 · H/C of the feedstock given in Tables
2.2 and 2.3 for coke and vacuum residue, respectively.
Table 3.1 lists theoretical maximal solar-to-chemical energy conversion efficien-
cies as a function of the cavity temperature for coke and vacuum residue and for
the 5 kW prototype reactor and its 300 kW scale-up. Typical values lie in the range
30 3. Experimental Setup

Table 3.1: Theoretical maximal solar-to-chemical energy conversion efficiency ηchem


for coke and vacuum residue for the 5 kW prototype reactor and its 300 kW scale-up
as a function of the cavity temperature.

Tcavity , K
feedstock Q̇solar , kW Daperture , mm 1300 1500 1700
ηchem , %
coke 5 50 64 56 48
VR 5 50 61 53 45
coke 300 500 62 52 41
VR 300 500 58 49 38

38–64 %. As expected, higher cavity temperatures lead to lower efficiencies as a con-


sequence of the increased reradiation losses. Values for vacuum residue are slightly
smaller that those for coke since more H2 bound in the feedstock has to be heated
(χVR > χcoke ). Finally, the scale-up reactor performs worse than the prototype re-
actor as a consequence of the lower concentration C at the aperture (C5kW ≈ 2500
suns, C300kW ≈ 1500 suns).

3.4.4 Residence Time


The average residence time of the reactants in the reactor, τ , is influenced by two
phenomena: (1) thermal expansion at the entrance of the reactor and (2) increase
of the molar flow due to the chemical reaction. Assuming a plug flow reactor and
first-order reaction the carbon conversion as a function of time t is given by [59]
X (t) = 1 − e−k·t , where k is the rate constant. Imposing the boundary conditions
X (0) = 0 and X (τ ) = XC it results:
ln(1−XC )
·t
X (t) = 1 − e τ (3.10)

where XC is the carbon conversion measured at the reactor’s outlet defined in


eq. (3.5). The total molar flux of gases ṅ (t) in the reactor is:

ṅ (t) = ṅAr + ṅH2 ,feedstock + ṅ0H2 O + ṅ0C · X (t) (3.11)

where ṅH2 ,feedstock accounts for the hydrogen bound in the coke released by pyrolysis
at the entrance of the reactor and the last term accounts for gas generated by
3.4. Key Process Parameters 31

the chemical reaction. Assuming ideal gas behaviour the fluid velocity is given by
dx/dt = u (t) = (ṅ (t) RT ) / (pS), where S is the reactor cross-section. Integration
with respect to time and imposing the boundary conditions x (0) = 0 and x (τ ) = L,
L being the length of the reactor, leads to:

p·V 1
τ= ·   (3.12)
R · Tcavity ṅ + ṅ 0 0 XC
Ar H2 ,feedstock + ṅH2 O + ṅC · 1 + ln(1−XC )

where V is the reactor volume, XC is the carbon conversion at the outlet, and all ṅi
are meant as initial feeding rates.
32 3. Experimental Setup
Chapter 4

Results

4.1 Coke — Dry Powder Feeding1


Operational conditions and a summary of measurements taken under approximate
steady-state conditions for 23 solar experimental runs are listed in Table 4.1. All
runs were performed with PD coke particles of type 1 with 2.21 µm mean initial
diameter. Reactants were continuously fed at a mass flow rate in the range of 2.0–
4.8 g coke/min and 6.0–9.0 g H2 O/min, and at ambient temperature and 423 K for
petcoke and steam, respectively. The average residence time of the reactants in the
reactor’s cavity, as calculated from eq. (3.12), varied between 0.7 and 1.3 s. Solar
power input through the aperture was in the range of 3.3–7.0 kW. The nominal
reactor temperatures varied between 1289 K and 1719 K. Chemical conversions for
petcoke and steam after a single pass, as defined by eqs (3.5) and (3.6), reached up
to 87% and 68%, respectively. Temperatures and gas compositions for a represen-
tative solar run (run 7) are shown in Fig. 4.1. In this run, the reactor was first
heated to above 1450 K under an argon flow. Thereafter, reactants were introduced
during an interval of 12 min at a rate of 3.5 g/min of coke and 6 g/min of H2 O,
corresponding to a H2 O/C molar ratio of 1.3. Average values at approximately
steady state condition are indicated by gray bars: Q̇solar =4.3 kW, Treactor = 1421
K, Tcavity =1201 K, ṅH2 =0.2 mol/min, ṅCO = 0.094 mol/min, ṅCO2 =0.04 mol/min,
and ṅCH4 =0.007 mol/min. The concentration of H2 S — a gasification byproduct
1
Material from this section has been published in: A. Z’Graggen, P. Haueter, D. Trommer, M.
Romero, J. C. de Jesus, and A. Steinfeld. Hydrogen production by steam-gasification of petroleum
coke using concentrated solar power — II. reactor design, testing, and modeling. International
Journal of Hydrogen Energy, 31:797–811, 2006.
34 4. Results

1600

temperature, K
Tcavity
product gas composition, % 1400

Tshell 1200
40
H2 1000
30
Q̇solar

Q̇solar , kW
20 5

CO 4
10 start of CO2
coke feeding 3
CH4
0
0 10 20 30
time, min

Figure 4.1: Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 7 in Table 4.1). The dashed vertical line indicates
the start of the petcoke feeding, the two vertical dotted lines delimit the interval
considered for the steady state calculations.

—, determined in separate measurements, reached up to 0.05% [101]. Its removal


may be accomplished downstream, for example by scrubbing. The syngas quality,
characterized by the molar ratios H2 /CO ≈ 2.2 and CO2 /CO ≈ 0.42, was in the
range of the values reported for conventional gasification [79].
The solar-to-chemical efficiency, ηchem , defined in eq. (3.7) ranges from 5% to
9%; the process efficiency, ηproc , defined in eq. (3.8) ranges from 22% to 35%. The
complete energy balance for each experimental run, ordered by increasing Tcavity , is
shown in Fig. 4.2. Indicated are the power lost by reflection and attenuation by
the window and reradiation from the cavity Q̇rerad , power lost by heat conduction
trough the reactor walls Q̇cond , power used to heat the reactants from 293 K to
Tcavity , Q̇heating , and change in enthalpy due to the chemical reaction Q̇chem (details
on the calculation are described in Chapter 6). Also given is the solar power input
Q̇solar and its estimated accuracy bounds (see also Appendix B). In the average,
heat losses are principally due to attenuation and emission by the window and re-
radiation through the aperture (≈ 17% ), and conduction through the reactor walls
(≈ 66%). As expected, the radiative losses are strongly temperature dependent. To
some extent they can be minimized by augmenting the input solar power flux, e.g.
4.1. Coke — Dry Powder Feeding 35

Q̇rerad Q̇cond Q̇chem Q̇heating


Q̇solar 1700
Ṫcavity

temperature, K
power, kW 6 1600
5
4 1500
3
1400
2
1 1300

experimental runs ordered by increasing Tcavity

Figure 4.2: Breakdown of the total input power into heating of the reactants Q̇heating ,
chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad and conduction losses
Q̇cond for the 23 runs of Campaign 1. Also plotted is the measured solar power
input, Q̇solar with its respective accuracy bounds.

by means of secondary optics — typically compound parabolic concentrators (CPC)


[114] — allowing the use of a smaller aperture for capturing the same amount of
energy. Increasing the reactor temperature further results in a higher reaction rate
and degree of chemical conversion, which in turn results in higher energy conversion
efficiencies.
36

ṁcoke ṁH2 O ṅAr ṅH2 ṅCO ṅCO2 ṅCH4 Q̇solar Tcavity Tshell XC XH2 O ηchem ηproc τ
# g/min mol/min kW K K – – – – s
1 2.3 6.0 0.31 0.22 0.09 0.046 0.004 4.5 1495 1276 0.83 0.55 0.07 0.29 1.0
2 2.0 6.0 0.23 0.17 0.07 0.035 0.005 5.0 1390 1193 0.75 0.42 0.05 0.22 1.2
3 2.4 6.0 0.27 0.24 0.10 0.046 0.008 4.6 1505 1263 0.87 0.58 0.07 0.31 1.0
4 2.1 6.0 0.27 0.18 0.08 0.038 0.008 4.4 1510 1203 0.78 0.46 0.06 0.27 1.0
5 2.3 6.0 0.27 0.12 0.04 0.026 0.011 3.3 1289 1067 0.47 0.28 0.05 0.23 1.3
6 2.2 6.0 0.27 0.15 0.06 0.031 0.007 4.3 1432 1099 0.62 0.38 0.05 0.23 1.1
7 3.5 6.0 0.27 0.20 0.09 0.040 0.007 4.3 1421 1210 0.54 0.52 0.07 0.26 1.1
8 2.9 6.0 0.27 0.23 0.11 0.041 0.006 5.3 1542 1221 0.74 0.58 0.07 0.27 1.0
9 3.5 6.0 0.27 0.28 0.13 0.044 0.008 6.1 1598 1208 0.71 0.66 0.07 0.28 0.9
10 3.6 6.0 0.27 0.29 0.14 0.043 0.004 5.7 1454 1201 0.71 0.68 0.07 0.28 1.0
11 3.9 6.0 0.27 0.27 0.13 0.040 0.007 4.6 1458 1227 0.63 0.64 0.09 0.31 1.0
12 4.5 6.0 0.27 0.29 0.15 0.038 0.004 6.7 1475 1219 0.59 0.68 0.06 0.25 1.0
13 4.5 6.0 0.27 0.28 0.13 0.037 0.008 5.9 1510 1208 0.54 0.62 0.07 0.26 0.9
campaign 1 conducted with dry petcoke particles.

14 3.4 7.0 0.27 0.26 0.12 0.047 0.007 5.6 1694 1214 0.71 0.56 0.07 0.29 0.8
15 3.5 8.0 0.27 0.26 0.13 0.049 0.007 5.8 1719 1224 0.72 0.50 0.07 0.29 0.7
16 3.5 9.0 0.27 0.26 0.12 0.053 0.008 6.6 1671 1231 0.70 0.45 0.06 0.26 0.7
17 3.7 9.0 0.27 0.30 0.12 0.057 0.010 5.3 1650 1230 0.70 0.48 0.08 0.32 0.7
18 3.9 9.0 0.27 0.31 0.14 0.056 0.012 5.7 1618 1220 0.71 0.50 0.08 0.32 0.7
19 4.8 9.0 0.27 0.33 0.15 0.057 0.011 7.0 1614 1213 0.61 0.53 0.07 0.28 0.7
20 3.5 9.0 0.27 0.32 0.13 0.064 0.009 4.6 1466 1266 0.78 0.51 0.09 0.35 0.8
21 3.5 9.0 0.27 0.29 0.12 0.058 0.011 5.0 1439 1223 0.75 0.48 0.08 0.32 0.8
22 3.5 9.0 0.27 0.23 0.09 0.047 0.014 4.3 1355 1137 0.59 0.37 0.08 0.29 0.9
23 2.9 9.0 0.27 0.26 0.11 0.055 0.004 4.6 1513 1325 0.79 0.44 0.08 0.32 0.8
Table 4.1: Steam and petcoke feeding rates, composition of the product gas, nom-
inal temperatures and performance parameters for 23 valid runs of experimental
4. Results
4.2. Coke — Slurry Feeding 37

4.2 Coke — Slurry Feeding2


Table 4.2 lists a summary of 29 experimental runs taken under approximate steady
state conditions. All runs were performed with PD coke. Three different particles
sizes were used: 1) ball milled, with an average particle diameter of 8.5 µm 2) sieved,
with particle diameters < 80 µm; and 3) sieved, with particle diameters < 200 µm,
corresponding to feedstock types 2, 3 and 4 of Table 2.1, respectively. The slurry
was prepared by stirring petcoke with demineralized water for a desired stoichio-
metric molar ratio (H2 O/C)slurry in the range 1–3. The total water-to-coke molar
ratio (H2 O/C)total , which includes the additional steam used to purge the window,
varied in the range 2.1–10.8. The petcoke mass flow rate was between 0.3 and 3.6
g/min. For better control purposes, an Ar flow of 0.13 mol/min was added to the
steam flow protecting the window. The resulting average residence time for the coke
particles was calculated using eq. (3.12), and τ was found to be in the range 0.9–2.3
seconds. Solar power input was varied over the range of 3.2–5.1 kW, resulting in
a nominal reactor temperature in the range of 1392–1566 K. Chemical conversion
for steam and petcoke after a single pass reached up to 28% and 87%, respectively.
The variation of temperatures and gas compositions during a representative solar
experimental run (run 10 of Table 4.2) is shown in Fig. 4.3. In this run, the reac-
tor was first heated to above 1500 K under an Argon flow. Thereafter, reactants
were continuously introduced during an interval of 14 minutes at a mass flow rate
of 2.9 g/min of slurry with (H2 O/C)slurry = 2. Additionally, 4 g/min of H2 O were
injected to protect the window, corresponding to (H2 O/C)total = 5.8. Main product
gases were H2 , CO, and CO2 . Relatively smaller amounts of CH4 were derived from
the pyrolysis of PD coke [102], believed to occur immediately after exposure to the
high-flux irradiation as a result of slurry heating rates exceeding 500 K/s. Average
values under approximate steady-state conditions, in the interval indicated by ver-
tical dashed bars of Fig. 4.3, were: Treactor =1536 K, ṅH2 =0.05 mol/min, ṅCO =0.02
mol/min, ṅCO2 =0.01 mol/min, and ṅCH4 =0.001 mol/min. H2 S concentration, de-
termined in separate measurements, reached up to 0.05% [101]; its removal may be
accomplished downstream, e.g., by scrubbing. Particle deposition or residual ashes
was not observed inside the reactor cavity. Previous thermogravimetric studies re-
2
Material from this section has been published in: A. Z’Graggen, P. Haueter, G. Maag, A.
Vidal, M. Romero, and A. Steinfeld. Hydrogen production by steam-gasification of petroleum coke
using concentrated solar power — III. reactor experimentation with slurry feeding. International
Journal of Hydrogen Energy, 32:992–996, 2007.
38 4. Results

1600

temperature, K
Tcavity
product gas composition, % 1400
25
Tshell H2 1200
20
1000
15 Q̇solar

Q̇solar , kW
5
10 start of CO
slurry CO2 4
5
feeding
CH4 3
0
0 5 10 15
time, min

Figure 4.3: Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 10 in Table 4.2). The dashed vertical line indicates
the start of the slurry feeding, the two vertical dotted lines delimit the interval
considered for the steady state calculations.

vealed a residuum of only 1.1 % weight after completion of the gasification reaction
[102]. The reactor’s thermal performance is described by the efficiencies ηchem and
ηproc defined in eq. (3.7) and eq. (3.8), respectively.
Efficiencies up to 5% and 23% were achieved for ηchem and ηproc , respectively.
An energy balance for all experimental runs is shown in Fig. 4.4. Also plotted is
the value of the solar power input Q̇solar . Main power losses are due to re-radiation
Q̇rerad and conduction Q̇cond , with average shares of 17% and 78%, respectively.
Q̇rerad , which takes into account attenuation by the spectrally selective window and
re-radiation through the aperture, was calculated using the radiosity method [124].
The power lost by conduction through the insulation Q̇cond was calculated assuming
a simplified cylindrical reactor geometry and using material properties provided
by the manufacturers. The power delivered for the chemical reaction Q̇chem was
calculated from the reaction enthalpy change at 298 K, whereas that stored in the
form of sensible heat of products was calculated from the species enthalpy change
between 473 K and the corresponding reaction temperature. Q̇steam accounts for the
evaporation and superheating of H2 O from 298 to 423 K (details are given in Chapter
6). Figure 4.5 shows the effect of particle size and slurry stoichiometry on the
4.2. Coke — Slurry Feeding 39

Q̇rerad Q̇cond Q̇chem Q̇heating


6
Q̇solar Ṫcavity
5 1600

temperature, K
power, kW
4
1500
3

2 1400

1
1300

experimental runs ordered by increasing Tcavity

Figure 4.4: Breakdown of the total input power into heating of the reactants Q̇heating ,
chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad and conduction losses
Q̇cond for the 29 runs of Campaign 2. Also plotted is the measured solar power
input, Q̇solar with its respective accuracy bounds.

ηchem ηproc XC
1
type 3
carbon conversion XC

type 4 25
0.8
20
efficiency, %
type 2

0.6
15
0.4
10
0.2 5

0 0
3:1 2:1 1.5:1 2:1 1.5:1 1:1
slurry molar ratio H2O/C

Figure 4.5: Carbon conversion and efficiencies for the 29 experimental runs of Cam-
paign 2 grouped by feedstock type and slurry molar ration H2 O/C. Feedstock types
2, 3 and 4 correspond to initial mean particle diameters of 6.7, 17.6 and 30.8 µm,
respectively.
40 4. Results

petcoke chemical conversion and energy conversion efficiency. Higher (H2 O/C)slurry
favors the reaction kinetics, but at the expense of higher mass flow rates and shorter
residence times, resulting in lower XC . It also results in a decrease of ηchem because
of the energy wasted to evaporate and superheat excess water. The opposite is
true for ηproc , assuming the sensible heat of excess steam exiting the reactor is
recovered. Best results were obtained for (H2 O/C)slurry = 1. Plugging of the feeding
systems was observed for (H2 O/C)slurry < 1 and 8.5 µm-particles. In principle, small
particles offer high specific surface area and, consequently, enhance the reaction rate.
Unreacted particles could be recycled in an industrial application.
D30 ṁcoke ṁH2 O (H2 O/C)slurry ṅAr ṅH2 ṅCO ṅCO2 ṅCH4 Q̇solar Tcavity Tshell XC XH2 O ηchem ηproc τ
# µm g/min mol/mol mol/min kW K K – – – – s
1 6.7 1.2 9.6 3.0 0.13 0.04 0.01 0.020 0.000 5.0 1566 1288 0.29 0.08 0.01 0.12 1.0
2 6.7 2.1 10.3 3.0 0.14 0.02 0.01 0.005 0.002 5.1 1516 1066 0.10 0.03 0.01 0.10 1.0
3 30.8 1.7 8.6 2.0 0.14 0.05 0.01 0.022 0.002 4.2 1514 1121 0.25 0.11 0.02 0.13 1.1
4 30.8 1.9 7.7 1.5 0.14 0.08 0.04 0.021 0.002 4.1 1478 1227 0.44 0.19 0.03 0.17 1.2
5 30.8 1.3 6.5 1.5 0.14 0.05 0.03 0.014 0.002 3.6 1431 1188 0.45 0.15 0.03 0.15 1.4
6 30.8 3.4 10.8 1.5 0.14 0.07 0.03 0.023 0.006 4.1 1393 1151 0.25 0.13 0.03 0.16 1.0
7 30.8 0.9 5.8 1.5 0.14 0.04 0.02 0.011 0.001 3.7 1530 1268 0.53 0.14 0.02 0.13 1.5
4.2. Coke — Slurry Feeding

8 30.8 0.3 4.7 1.5 0.13 0.01 0.00 0.004 0.000 3.3 1542 1275 0.31 0.04 0.00 0.08 1.8
9 30.8 1.9 7.7 1.5 0.13 0.08 0.04 0.021 0.002 3.7 1533 1288 0.42 0.18 0.03 0.18 1.2
10 17.6 0.8 6.1 2.0 0.13 0.05 0.02 0.011 0.001 3.8 1536 1283 0.51 0.12 0.02 0.14 1.5
11 17.6 1.8 8.7 2.0 0.13 0.10 0.04 0.021 0.002 4.0 1511 1253 0.50 0.17 0.04 0.21 1.1
12 17.6 1.1 6.1 1.5 0.13 0.06 0.02 0.014 0.001 4.3 1540 1295 0.44 0.14 0.02 0.13 1.4
13 17.6 2.0 8.1 1.5 0.13 0.11 0.04 0.025 0.003 4.7 1507 1264 0.45 0.20 0.03 0.18 1.1
14 17.6 3.6 11.2 1.5 0.13 0.16 0.06 0.034 0.005 4.7 1489 1244 0.38 0.21 0.05 0.23 0.9
15 17.6 1.4 6.8 1.5 0.16 0.08 0.03 0.018 0.002 4.6 1466 1195 0.49 0.18 0.02 0.14 1.3
campaign 2 conducted with a coke-water slurry.

16 17.6 1.4 4.8 1.5 0.16 0.08 0.03 0.016 0.002 4.7 1475 1209 0.51 0.25 0.02 0.14 1.6
17 17.6 1.4 5.8 1.5 0.16 0.09 0.04 0.019 0.002 4.8 1484 1220 0.60 0.24 0.03 0.15 1.4
18 17.6 0.6 4.8 1.0 0.13 0.04 0.01 0.009 0.001 3.2 1547 1309 0.49 0.11 0.01 0.14 1.7
19 17.6 2.4 7.2 1.0 0.13 0.11 0.05 0.020 0.003 3.8 1510 1249 0.38 0.21 0.04 0.20 1.2
20 17.6 1.4 5.8 1.0 0.13 0.08 0.03 0.015 0.002 4.1 1502 1234 0.52 0.20 0.03 0.16 1.5
21 17.6 1.4 4.8 1.0 0.13 0.08 0.04 0.015 0.001 4.0 1533 1266 0.53 0.25 0.03 0.16 1.6
22 17.6 1.4 3.8 1.0 0.13 0.07 0.03 0.013 0.001 3.9 1543 1276 0.48 0.28 0.03 0.15 1.8
23 17.6 0.6 4.8 1.0 0.13 0.04 0.02 0.009 0.000 3.6 1496 1245 0.69 0.15 0.02 0.13 1.7
24 17.6 0.6 4.8 1.0 0.09 0.05 0.02 0.011 0.001 3.7 1514 1257 0.80 0.17 0.02 0.14 1.9
25 17.6 0.6 3.8 1.0 0.09 0.06 0.03 0.012 0.002 3.7 1518 1262 0.87 0.23 0.02 0.14 2.2
26 17.6 0.6 2.8 1.0 0.14 0.05 0.02 0.009 0.001 3.8 1525 1269 0.73 0.26 0.02 0.12 2.3
27 17.6 1.4 5.8 1.0 0.09 0.08 0.03 0.014 0.002 3.8 1512 1258 0.47 0.18 0.03 0.16 1.6
28 17.6 1.4 4.8 1.0 0.09 0.07 0.03 0.012 0.002 3.8 1499 1244 0.44 0.21 0.03 0.15 1.8
29 17.6 2.4 9.2 1.0 0.09 0.13 0.05 0.024 0.003 4.4 1455 1198 0.44 0.20 0.04 0.21 1.1
Table 4.2: Water and petcoke feeding rates, composition of the product gas, nom-
inal temperatures and performance parameters for 29 valid runs of experimental
41
42 4. Results

3
4.3 Vacuum Residue — Liquefied Injection
Table 4.3 lists the operational parameters and results of 12 solar experimental runs
taken under approximate steady-state conditions. Mass flow rate of VR ṁVR was
in the range 0.65–4.17 g/min. Mass flow rate of steam ṁH2 O — injected coaxially
— was in the range 2–6 g/min, corresponding to a H2 O/C molar ratio between 1.12
and 7.18. Ar was injected through purging nozzles to protect the window at a mass
flow rate in the range 0.27–0.36 mol/min. The resulting average residence time for
the coke particles was calculated using eq. (3.12), where ṅH2 ,feedstock denotes the
hydrogen bound in the VR, which is released by pyrolysis at the entrance of the
reactor. Residence time τ for all runs was in the range 0.8–1.8 seconds. The solar
power input Q̇solar varied in the range 3.6–6.8 kW, which corresponds to a mean solar
flux concentration ratio over the reactor’s aperture in the range 1833–3463 suns (1
sun = 1 kW/m2 ). The nominal reactor temperature Treactor was in the range 1420–
1567 K. Typical heating rates exceeded 500 K/s. VR was firstly pyrolyzed, releasing
volatile hydrocarbons such as C2 H2 , C2 H4 , and C2 H6 as reported previously [73, 6].
The remaining solid particles, typically about 15–24 %wt of the original VR —
denoted Conradson carbon residue (CCR, [30]) — underwent steam-gasification to
produce primarily H2 , CO, and CO2 , consistent with the thermodynamic equilibrium
predictions. Chemical conversion for steam XH2 O and carbon XC reached up to 44%
and 50%, respectively.
The variation of temperatures and gas compositions during a representative so-
lar experimental run (run 12 of Table 4.3) is shown in Fig. 4.6. In this run, the
reactor was first heated to above 1550 K under an Ar flow. Thereafter, reactants
were continuously introduced during an interval of 10 minutes at ṁVR =1.2 g/min
and ṁH2 O =2 g/min, corresponding to H2 O/C molar ratio of 1.3. Average molar
flow rates of gaseous products, measured at Treactor =1567 K under approximately
steady-state conditions in the interval indicated by vertical dashed bars of Fig. 4.6,
were: ṅH2 =0.08 mol/min, ṅCO =0.03 mol/min, ṅCO2 =0.01 mol/min, and ṅCH4 =0.002
mol/min. Negligible amounts of volatile hydrocarbons (C2 H4 and C2 H2 at 11 and
60 ppm, respectively) were derived from the pyrolysis of VR, occurring immediately
after exposure to the high-flux solar irradiation. Approximately 40%wt of the CCR
3
Material from this section has been published in: A. Z’Graggen, P. Haueter, G. Maag, M.
Romero, and A. Steinfeld. Hydrogen production by steam-gasification of carbonaceous materials
using concentrated solar energy — IV. reactor experimentation with vacuum residue. International
Journal of Hydrogen Energy,33:679–684, 2008.
4.3. Vacuum Residue — Liquefied Injection 43

Tcavity 1600

temperature, K
product gas composition, %
Tshell 1400

1200
15
H2 1000
10 Q̇solar

Q̇solar , kW
start of 5
CO
5 VR 4
feeding CO2
CH4 3
0
0 2 4 6 8 10 12
time, min

Figure 4.6: Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 12 in Table 4.3). The dashed vertical line indicates
the start of the petcoke feeding, the two vertical dotted lines delimit the interval
considered for the steady state calculations.

Q̇rerad Q̇cond Q̇chem Q̇heating


Q̇solar
1600
Ṫcavity
temperature, K

6
power, kW

5 1500
4
3 1400
2
1 1300

experimental runs ordered by increasing Tcavity

Figure 4.7: Breakdown of the total input power into heating of the reactants Q̇heating ,
chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad and conduction losses
Q̇cond for the 12 runs of Campaign 3. Also plotted is the measured solar power
input, Q̇solar with its respective accuracy bounds.
44 4. Results

χH2 ,gasification ṁH2 O

relative origin of product H2


χH2 ,pyrolysis ṁVR
1 7
6
0.8

ṁ, g/min
5
0.6 4
0.4 3
2
0.2
1
0 0
experimental runs ordered by increasing ṁVR

Figure 4.8: Fractions of hydrogen in the product gas derived by either pyrolysis or
gasification, and the feeding rates of reactants VR and steam for experimental runs
of Campaign 3.

deposited as solid agglomerates inside the cavity (size up to 5 mm), which contained
90% carbon, 3.5% sulfur, and 5% metallic fraction constituted by 75% V and 18%
Ni. This composition is comparable to that of Petrozuata Delayed coke [101]. The
rest consisted of particles with a carbon content of 96.5%, a BET specific surface
area of 20.5 m2 /g (determined with Micromeritics TriStar Analyzer), and a mean
particle size of 6.5 µm(determined by laser scattering with a Horiba LA950). Energy
conversion efficiencies up to 2% and 17% were achieved for ηchem and ηproc , respec-
tively. Main heat losses are due to re-radiation through the aperture and conduction
through the insulation, with average shares of 17% and 73%, respectively, as shown
in Fig. 4.7. Re-radiation losses, which take into account attenuation by the spec-
trally selective quartz window and emission through the aperture, were calculated
using the radiosity method [124]. Conduction heat losses, which take into account
heat bridges through the oil-cooled components, were calculated using the thermal
conductivities of the insulation materials, as provided by the manufacturer (for de-
tails see Chapter 6). Optimization for minimizing heat losses and maximizing the
solar energy conversion efficiency was outside the scope of this study. Figure 4.8
reveals the origin of the H2 in the product gas. Plotted for all experimental runs are
the fractions of hydrogen derived by either pyrolysis or gasification, χH2 ,gasification and
χH2 ,pyrolysis , and the feeding rates of reactants, ṁVR and ṁH2 O (ordered by increasing
4.3. Vacuum Residue — Liquefied Injection 45

ṁVR ). Average values were χH2 ,pyrolysis =63% and χH2 ,gasification =37%. As expected,
a higher ṁVR resulted in a higher χH2 ,pyrolysis since H2 contained in the VR is com-
pletely released during the pyrolysis step. In contrast, ṁH2 O had negligible effect
because H2 O was fed in excess and only partly consumed by the gasification process.
Ideally, complete conversion of carbon bound in the vacuum residue would result in
χH2 ,pyrolysis =40% and χH2 ,gasification =60%.
46

ṁVR ṁH2 O ṅAr ṅH2 ṅCO ṅCO2 ṅCH4 CC2 H4 CC2 H6 CC2 H2 Q̇solar Tcavity Tshell XC XH2 O ηchem ηproc τ
# g/min mol/min ppm kW K K – – – – s
1 1.0 3.0 0.29 0.05 0.02 0.010 0.005 1462 9 2238 5.6 1420 1096 0.44 0.23 0.01 0.09 1.5
2 4.2 6.0 0.36 0.17 0.03 0.008 0.013 1467 29 1943 5.6 1488 1160 0.16 0.13 0.02 0.18 0.8
3 0.7 6.0 0.36 0.05 0.01 0.011 0.002 32 0 214 4.2 1472 1232 0.50 0.10 0.01 0.13 1.0
paign 3 conducted with liquefied VR.

4 1.2 6.0 0.36 0.07 0.01 0.010 0.005 88 0 619 3.8 1552 1274 0.31 0.10 0.02 0.17 1.0
5 2.2 6.0 0.36 0.11 0.02 0.010 0.008 308 2 1381 4.1 1543 1271 0.24 0.12 0.02 0.19 0.9
6 2.2 6.0 0.27 0.11 0.02 0.013 0.005 231 2 999 5.4 1534 1233 0.27 0.15 0.02 0.15 1.0
7 0.7 4.0 0.27 0.04 0.01 0.008 0.002 53 0 344 3.9 1491 1264 0.41 0.12 0.01 0.11 1.4
8 1.2 4.0 0.27 0.06 0.01 0.008 0.003 80 0 570 3.6 1501 1298 0.27 0.13 0.01 0.14 1.3
9 1.7 4.0 0.27 0.08 0.02 0.009 0.004 88 1 594 4.2 1540 1290 0.24 0.16 0.02 0.14 1.2
10 0.7 2.0 0.27 0.03 0.01 0.007 0.001 16 0 124 4.1 1503 1320 0.34 0.20 0.01 0.08 1.8
11 1.2 2.0 0.27 0.05 0.01 0.007 0.005 408 3 1345 6.8 1513 1191 0.32 0.27 0.01 0.07 1.7
12 1.2 2.0 0.28 0.08 0.03 0.010 0.002 11 0 60 4.9 1567 1332 0.48 0.44 0.02 0.12 1.5
temperatures and performance parameters for 12 valid runs of experimental cam-
Table 4.3: Steam and VR feeding rates, composition of the product gas, nominal
4. Results
4.4. Comparison 47

4.4 Comparison

Table 4.4: Average operational parameters and results of the solar experimental
campaigns conducted with dry coke powder, cokewater slurry, and VR.

feedstock dry coke powder4 coke-water slurry VR4


ṁfeedstock , g/min 3.3 1.47 1.50
ṁH2 O , g/min 7.17 6.61 4.25
(H2 O/C), mol/mol 1.71 4.0 2.74
Tcavity , K 1514 1506 1510
XC , % 68.9 47.5 33.1
XH2 O , % 52.6 17.3 17.6
Q̇solar , W 5177 4079 4683
Xchem , % 7.0 2.5 1.4
Xproc , % 28.3 15.2 13.2
τ, s 0.92 1.40 1.27
4
Steam fed separately

Table 4.4 summarizes the main average process parameters for three experimental
campaigns performed with comparable solar power inputs and reactor temperatures,
and with three different feedstocks: (1) dry coke powder (with steam fed separately),
(2) coke-water slurry, and (3) VR (with steam fed separately). The campaign with
dry coke powder outperforms the other two in terms of chemical conversion and en-
ergy efficiency. For example, ηproc is 28.3%, 15.2% and 12.3% for dry coke powder,
slurry, and VR, respectively. In the slurry campaign, the overall performance was
diminished by excess water fed as a result of the high H2 O/C molar ratios required
for the slurry. In the VR campaign, besides using excess water, the formation of
solid agglomerates affected negatively the chemical conversion. The implementation
of a spray nozzle may facilitate dispersion into smaller particles and, consequently,
enhance the reaction kinetics. The aerodynamic protection of the window worked
well for the duration of all experimental runs, each about 45 minutes. The win-
dow was maintained clean and clear from particle deposition or condensable gases.
The oil-cooled frustum served as a buffer zone between the window and the cavity
(reaction chamber), and prevented steam condensation. In a scale-up application
for a solar tower concentrating facility, this annular component may be heated with
spilled radiation from the heliostat field. Argon should be completely substituted
by steam to avoid the energy and cost of recycling an inert gas.
48 4. Results
Part II

Modeling
Chapter 5

Simulation Framework

The simulations presented in the second part of this thesis share a common set
of boundary conditions and models for the chemical kinetics and the polydisperse
particulate medium. The characteristics of the incoming solar radiation as well as the
conductive behavior of the reactor walls were determined by stand-alone preliminary
simulations. Wall and window radiative properties are presented. The concept of
‘equivalent monodisperse diameter’ is introduced to model the polydisperse feedstock
types, which have random size-distribution functions. Finally, the kinetic rate laws
for steam gasification of petcoke developed by Trommer [102] are presented.

5.1 General Boundary Conditions


5.1.1 Incoming Radiation
An experimentally validated 3D Monte-Carlo ray-tracer was used the determine the
exact properties of the concentrated radiation incident on the reactor’s aperture.
The simulation setup is shown in Fig. 5.1. The location and direction of incident
rays on the aperture was recorded taking into account the geometrical characteristics
of the concentrating facilities, specular reflection errors, and non-parallelism of the
sun rays. Figure 5.2 shows the solar power through the aperture Q̇solar and the
solar flux q̇solar as a function of the aperture’s radius for (a) PSI’s solar furnace
(nominal Q̇solar =5 kW), and (b) CIEMAT’s solar tower (nominal Q̇solar =300 kW,
data from CIEMAT). Shown in 5.3 is the averaged angular solar flux distribution at
the aperture, used as input to the solar reactor. Finally, the solar spectrum of the
incoming radiation is approximated by Planck’s blackbody emission at 5780 K.
52 5. Simulation Framework

incident solar reactor


radiation aperture

10
concentrated
5 radiation

z, m 0

{5 reactor

{10 heliostat
0
parabolic concentrator 10
10
20 0
x, m y, m
30
{10

Figure 5.1: Simulation setup used to determine the angular and radial distributions
of the incoming concentrated solar radiation at PSI’s solar furnace [34].

The same simulation setup was used in the early stages of the engineering design
to estimate the radiative flux intensities to be expected inside the cavity. Figure 5.4
shows lines of equal incident solar flux q̇solar in kW, projected on planes parallel to
the aperture. An aperture diameter of 50 mm and a nominal power through the
aperture of 5 kW were considered, while no interaction with the cavity walls was
modeled. Maximal q̇solar >3050 W/m2 is reached on a small spot at the center of
the aperture. Peak values of 2959, 2815, 2088 and 1318 W/m2 are predicted for
planes located 1, 2, 3 and 4 cm behind the aperture. Also indicated is the cavity of
the prototype reactor presented in Chapter 3 (gray). In that reactor particles are
injected on a plane 4 cm away from the aperture in order to take advantage of the
still highly concentrated radiation (1318 kW/m2 ), on the one hand, and to prevent
the particles to exit the reactor through the aperture on the other.
5.1. General Boundary Conditions 53

(a) (b)

2.5
3 5 300

2 250
q̇solar , MW/m2

q̇solar , MW/m2
4

Q̇solar , kW

Q̇solar , kW
2 200
1.5
3
150
2 1
1 100
1 0.5
50

0 0 0 0
0 0.025 0.05 0 0.25 0.5
Daperture , m Daperture , m

Figure 5.2: Solar radiative flux (left axis) and solar power (right axis) as a function
of the aperture’s diameter for (a) PSI’s solar furnace, and (b) CIEMAT’s solar tower.

0
−π/8 π/8
−π/4 π/4

PSI’s solar incident


furnace
CIEMAT’s radiation
solar tower

Figure 5.3: Angular distribution of the concentrated solar radiation at the aperture
for PSI’s solar furnace and CIEMAT’s solar tower.
54 5. Simulation Framework

100 0.05
50 mm

50
aperture

0
0.025

100
280
0

0
20
305

00

r, m
0 0
incoming

250
radiation -0.025

-0.05
0 0.02 0.04 0.06 0.08 0.1
x, m

Figure 5.4: Lines of equal projected incident solar flux q̇solar in kW/m2 on planes
parallel to the aperture. Results are given for a 50 mm aperture and a nominal
power trough the aperture Q̇solar =5 kW. Also indicated is the shape of the prototype
reactor’s cavity (gray).

5.1.2 Reactor Walls1


Figure 5.5 shows the heat fluxes at a cavity wall element considered in the simulation.
Conductive heat transfer was preliminarily simulated with a 2D axi-symmetric FE
solver. The exact reactor geometry shown in Fig. 3.1 was considered, including the
inner ceramic Al2 O3 liner (k = 25 W/mK), the Inconel 601 shell (k = 11.2 W/mK),
the outer insulation (Insulform 160, k=0.09–0.3 W/mK), the oil-cooled aluminium
frustum (k = 186 W/mK) as well as the graphite insulation between the shell and
the frustum and the 1 mm gap between the Al2 O3 liner and the shell. Typical
radiative loads on the liner were calculated with the raytracer presented in Section
5.1.1 and for measured Q̇solar , the heat removed by the oil cooling was estimated
from the measured temperature differences between inlet and outlet. Finally, natural
convection for a cylinder was assumed on the outside. Figure 5.6 shows the average
overall heat transfer coefficient extracted from 23 simulation runs based on the
measurements of Campaign 1. Values range between 12 W/ (m2 K) at the center of
the cavity to 180 W/ (m2 K) close to the aperture due to heat bridges. The overall
1
The FE simulation used in this section has been set up in the framework of: L. Donati,
Hydrogen production by steam-gasification of petroleum coke using concentrated solar power —
Thermal Simulation of the Synpet Solar Reactor with ANSYS, Semester Thesis, ETH Zurich, 2005.
5.1. General Boundary Conditions 55

outside inside

q̇cond q̇ic

4
σTwall
T1

natural Twall
convection Tshell
reactor insulation reactor shell (Inconel)
(Al2O3 / ZrO2)
reactor liner (Al2O3)

Figure 5.5: Schematic of heat fluxes at the wall. Radiation emitted by particles
and other cavity wall elements, q̇ic , is incident from the inside (right), while the wall
itself emits radiation toward the inside. On the outside (left) the reactor is cooled
by natural convection. Finally, conduction occurs through the three material layers
of the wall.

heat transfer coefficient U is related to the oil temperature on the front cone and to
the ambient temperature T∞ for the rest of the cavity.

Diffuse gray surfaces ( = 0.8) are assumed for the reactor’s inner walls because
of deposited petcoke particles. The impact of the uncertainty of  on Tcavity and
XC was assessed by sensitivity analysis (see Section 7.4). In the average, a change
of 10% in  induced a change of 0.4 % in Tcavity and XC . Convective heat transfer
from the walls to the fluid is found by CFD to be two orders of magnitude smaller
than the heat transfer by radiation and conduction, and is therefore neglected. An
energy balance at the wall leads then to the following implicit expression for the
temperature: Z ∞
4
q̇ic,λ λ dλ − σTwall − q̇cond = 0 (5.1)
0

where q̇ic is the incident thermal radiation, the second term accounts for radiation
emitted from the wall, and the third term accounts for the conduction heat loss
through the wall:
q̇cond = U · (T∞ − Twall ) (5.2)
56 5. Simulation Framework

160

U , W/(m2 K) 120
front cone
80 rear cone
cylinder

40

0
0 5 10 15 20 25
x, cm

Figure 5.6: Overall heat transfer coefficient U for conductive losses trough the reactor
walls as a function of the location along the reactor axis. Also indicated is the
longitudinal cross-section of the reactor’s cavity.

5.1.3 Quartz Window

The spectral directional reflectivity ρλ (θ) of the quartz window surface is calculated
with the Fresnel equations [68] and the complex refractive index n + ik from [72],
while the absorption coefficient of the window is given by κλ = 4πkλ /λ. The ‘one
pass’ transmittance trough the window τ is then given by τ = e−κ·s , where s is
the distance traveled in the quartz, determined by the window thickness and the
incidence angle θ. Assuming similar refractive indexes on both sides of the win-
dow, the total transmittance, reflectance and absorptance used in the simulation are
expressed by (subscripts λ omitted for clarity) [68]:

(1 − ρ)2 τ
Tr w = (5.3)
1 − ρ2 τ 2
!
(1 − ρ)2 τ 2
Rw = ρ 1+ (5.4)
1 − ρ2 τ 2
(1 − ρ) (1 − τ )
Aw = (5.5)
1 − ρτ

Figure 5.7 shows the complex refractive index of quartz and the total transmittance
Tr w of a 3 mm thick window for selected incident angles θ = 0, 45 and 60◦ as a
5.1. General Boundary Conditions 57

Tr w,0 Tr w,45 1
2.5
Tr w,60 0.8
2
n 0.6

Tr w,θ
n; k

1.5
0.4
1

0.5 0.2
k
0 0
0.1 1 10
λ, µm

Figure 5.7: Complex refractive index of quartz n + ik [72] and transmittance Tr w


for a 3 mm thick window and for three selected incident directions θ as a function
of wavelength λ.

function of the wavelength λ. Strong selective behavior of the window is observed,


with virtually zero transmittance below 0.2 µm and above 4 µm and values higher
than 0.8 in between. The influence of the incidence angle is marginal; for example
average values of 0.93, 0.91 and 0.83 are found for the normal direction (0◦ ), for the
maximal acceptance angle of the facility (45◦ ) and for 60◦ , respectively.

Figure 5.8 shows the heat fluxes considered at the window. The window temper-
ature is assumed constant over it’s thickness of 3 mm. Convective heat losses on the
outside are then given by q̇conv = h · (T∞ − Tw ) where h = 263 W/(m2 K) is obtained
from the empirical correlation Nux = 0.0308 · Re4/5 x · Pr
1/3
for turbulent flow over
a flat plate [44], as the window is cooled externally by a ventilator. The impact
of the uncertainty of h on Tcavity and XC was assessed by sensitivity analysis (see
Section 7.4). In the average, a change of 10% in h induced a negligible change of
0.03 % in Tcavity and XC . Similarly to eq. (5.1) heat balance for the window yields:
Z ∞
(q̇ic,λ + q̇solar,λ ) Aw,λ dλ − 2σAw (Tw ) Tw4 − q̇conv = 0 (5.6)
0
58 5. Simulation Framework

outside inside
µ = 45º
Tw
q̇solar q̇ic

σAw Tw4 σAw Tw4

T1
q̇conv window

ventilator

Figure 5.8: Schematic of heat fluxes at the window. Concentrated solar radiation
q̇solar is incident from the outside of the reactor (left) whereas radiation emitted by
the cavity walls and particles, q̇ic , is coming from the inside (right). The window itself
emits radiation on both sides. Finally, the window is cooled by forced convection on
the outside.

5.2 Polydisperse Coke Particles


The polydisperse medium is characterized by its population density, i.e. the particle
size distribution function f describing the number of particles found in an infinites-
imal interval around diameter D. For a sample containing a discrete number of
particles Np with respective diameters Di the continuous population density f is
given by the sum of each particle’s contribution:

Np Np
1 X 1 X
f (D) = δ (D − Di ) = w (D − Di , ∆D) (5.7)
Np i=1 Np i=1

where the Dirac delta function δ is substituted for a symmetric kernel function w
R∞
around 0 with bandwidth ∆D and with 0 w dD = 1 (typically, w is a Gaussian
error distribution curve). The solid lines in Figure 5.9 show f (D) · D3 , the volume
density measured by laser scattering, for feedstocks type 1 and 3 considered here
(see also Fig. 2.2). As a consequence of the chemical reaction the particles shrink by
a factor ηsk = Di (XC > 0) /Di (XC = 0), thus influencing the distribution function:

Np
1 X
w D − Di0 · ηsk , ∆D

fsk (D) = (5.8)
Np i=1
5.2. Polydisperse Coke Particles 59

XC = 0.0
XC = 0.5
XC = 0.75
type 3
type 1
f (D) · D3

1 10 100
D, µm

Figure 5.9: Volume density of the polydisperse particles used in Campaigns 1 (type
1) and 2 (type 3). Solid lines are measured values for the unreacted samples, dashed
and dashed-dotted lines are calculated results at carbon conversions of 0.5 and 0.75,
respectively.

where ηsk is a function of the particle diameter and of the extent of the reaction (see
Section 5.3 for details about the differences in reactivity as a function of particle
size).
Two examples of the modified distribution function are given in Fig. 5.9, for a
total reaction extent XC =0.5 and 0.75:
R∞
0
fsk · (D · ηsk )3 dD
XC = 1 − R∞ (5.9)
0
f · D3 dD

As expected, bigger particles tend to react slower than the small ones leading to a
shape change of the distribution function as the chemical reaction progress, observ-
able especially for feedstock type 3. For simplification, the polydisperse medium can
be described by a set of equivalent monodisperse diameters defined generally as:

 1
f (D) Dp dD ( p−q )
R ∞
0
Dpq = R∞ (5.10)
0
f (D) Dq dD

where p and q are integer numbers between 0 and 3. D30 , D31 , D32 are applied for
60 5. Simulation Framework

100
type 3

10
D, µm

1

D32
D31 type 1
0.1 D30
0 0.75 0.95 0.99 1
XC

Figure 5.10: Equivalent monodisperse diameters calculated with eqs (5.10) and
(5.26) of the polydisperse particles used in Campaigns 1 (type 1) and 2 (type 3)
as a function of the carbon conversion.

volume-based, volume-to-diameter-ratio based, and volume-to-surface-ratio based


phenomena, respectively (see also eqs (5.13) and (7.13) for examples of application).
Figure 5.10 shows the variation of the equivalent monodisperse diameters as a func-
tion of the reaction extent XC for feedstock types 1 and 3 with mean initial sizes D30
of 2.21 and 17.58 µm, respectively. The difference between the equivalent diameters
indicate the importance of considering the polydisperse nature of the particulate
medium, especially for the feedstock type 3, where the difference reaches one order
of magnitude. Note that eqs (5.10) and (5.9) lead to a simple relation for the mean
diameter of the polydisperse medium:
1
0
D30 = D30 · (1 − XC ) 3 (5.11)

5.2.1 Radiative Properties


Values for the complex refractive index of coke are given in Table 5.1 [14, 29]. For the
solid volume fractions fV < 10−3 considered in this study the independent scattering
regime is valid [91]. Continuum absorption, scattering, and extinction coefficients
for the polydisperse medium are then found as a function of the corresponding
5.2. Polydisperse Coke Particles 61

Table 5.1: Complex refractive index of coke [14, 29].

λ, µm n k
0.308 1.96 0.869
1.35 2.04 0.901
2.40 2.11 0.934
3.44 2.19 0.966
4.49 2.27 0.998
5.53 2.35 1.03

properties of single spheres [68]:


R∞
3fV 0
{Qa,λ , Qs,λ , Qext,λ } f (D) D2 dD
{κλ , σsλ , βλ } = · R∞ (5.12)
2 0
f (D) D3 dD

For size parameter ξ = πD/λ in the range 0.1 − 103 considered here, Mie theory
applies [11]. The efficiency factors for absorption, scattering and extinction of a
single sphere, Qa,λ , Qs,λ , and Qext,λ are calculated based on the complex refractive
index of the solid material using the routine BHMIE [11]. For ξ > 5, geometrical
optics applies and eq. (5.12) simplifies to:

3 fV
{κλ , σsλ , βλ } = · {Qa,λ , Qs,λ , Qext,λ } (5.13)
2 D32

Comparison of the values calculated with eq. (5.12) and with the simplified approach
of eq. (5.13) showed negligible differences for the particle size distributions considered
in this study (note that for the comparison Qa,λ , Qs,λ , and Qext,λ in eq. 5.13 were
evaluated at D32 ). κλ , σsλ and βλ of the polydisperse particle cloud are therefore
calculated by means of eq. (5.13) based on the equivalent diameter for radiation,
D32 .

Figures 5.11 (a) and (b) show the spectral absorption and scattering coefficients
of the particulate medium in Campaigns 1 (feedstock type 1) and 2 (type 3), respec-
tively, for a typical initial volume fraction fV = 5 · 10−5 , and for carbon conversions
XC = 0, 0.5, and 0.75. The smaller particles of Campaign 1 absorb more than 10
times better that those of Campaign 2.

Analogous to eq. (5.12) the scattering phase function of the polydisperse particle
62 5. Simulation Framework

(a) (b)
120 8

XC = 0.0 XC = 0.0
80
κ λ ; σλ,s , 1/m

κ λ ; σλ,s , 1/m
4

0.5 0.5
2
40 0.75
0.75
1
σλ,s σλ,s
20 κλ κλ
0.5 1 2 4 0.5 1 2 4
λ, µm λ, µm

Figure 5.11: Spectral distribution of the absorption and scattering coefficients, cal-
culated for a typical volume fraction of 5 · 10−5 and carbon conversions of 0.0, 0.5
and 0.75, for the feedstock used in (a) Campaign 1, and (b) Campaign 2.

cloud is found by a weighted sum of the phase functions of each particle size [68]:
R∞
0
Φλ (D, ω0 ) · Qs,λ · f (D) D2 dD
ΦTλ (ω0 ) = R∞ (5.14)
0
Qs,λ · f (D) D2 dD

where Φλ is calculated with the routine BHMIE [11]. Because of computational time
restraints, the Henyey-Greenstein approximation [36] is introduced,

2
1 − gTλ
ΦTλ (D, θ0 ) = 3/2
(5.15)
2
(1 + gTλ − 2gTλ · cos θ0 )

where the asymmetry factor gTλ is calculated similarly to eq. (5.14):


R∞
0
gλ (D) · Qs,λ · f (D) D2 dD
gTλ = R∞ (5.16)
0
Qs,λ · f (D) D2 dD

and the asymmetry factor gλ for a single particle size is defined as


1
R
gλ = 4π Φ (D, θ) sin θ dΩ.
4π λ
The scattering phase functions for particles of Campaign 1 (feedstock type 1)
are shown in Figs 5.12 (a) and (b) for carbon conversions 0.0 and 0.75, respec-
5.2. Polydisperse Coke Particles 63

(a)
Mie for D32 = 3.134 µm
10
3 Mie, polydisperse at XC = 0.0
HG, polydisperse at XC = 0.0
2
10
λ = 0.5 µm
Φ
1 λ = 2.0 µm
10

0
10

−1
10
0 π/4 π/2
θ, rad

(b)
Mie for D32 = 1.99 µm
103
Mie, polydisperse at XC = 0.75
HG, polydisperse at XC = 0.75

102 λ = 0.5 µm
Φ

101 λ = 2.0 µm

100

10−1
0 π/4 π/2
θ, rad

Figure 5.12: Scattering phase function for the equivalent diameter D32 , for the
polydisperse medium (eq. (5.14)), and for the Henyey-Greenstein approximation
(eq. (5.15)), calculated for carbon conversions of (a) 0.0; and (b) 0.75, for the
feedstock used in Campaign 1. Directions θ = π/2 – π not shown in plot, because
no backward scattering peak was observed.

tively. Two selected wavelengths λ=0.5 and 2 µm are considered, corresponding to


the locus of maximum emissive power Eλb for radiation emitted at 5780 and 1450
K, respectively. Curves are plotted for a monodisperse particulate medium with
equivalent diameter D32 , for the polydisperse particulate medium of Fig. 5.9, and
64 5. Simulation Framework

for the Henyey-Greenstein (HG) approximation (eq. (5.15)). Preference for forward
scattering is observed. As expected, the polydisperse particle cloud has a smoother
scattering phase function, which is well approximated by the HG equation.

5.3 Chemical Kinetics

Table 5.2: Arrhenius parameters of the kinetic rate constants for the steam gasifi-
cation of coke [100].

EA , J/mol k0 , 1/s
K1 2.707 · 105 1.158 · 103
K2 1.615 · 102 4.978 · 10−1
K3 −9.404 · 104 1.149 · 10−7
K4 7.068 · 103 4.033 · 10−8
K5 4.551 · 102 8.152 · 10−6

The chemical kinetics of PD coke steam-gasification were extensively investigated


by D. Trommer [102]. This section summarizes the findings pertinent to the present
study. The overall chemical conversion can be represented by the simplified net
reaction: z 
CHz Oy + (1 − y) H2 O = + 1 − y H2 + CO (5.17)
2
where z and y are the elemental molar ratios of H/C and O/C in the coke, re-
spectively. The enthalpy change of the endothermic reaction is ∆HR |298K = 131
kJ/molC . Table 2.2 lists the approximate main elemental chemical composition, the
low heating value (LHV), and elemental molar ratios of H/C and O/C for Petrozu-
ata Delayed (PD) coke used in the experimental campaign. The gasification kinetic
model is based on the oxygen-exchange mechanism describing reversible O-transfer
surface reactions followed by an unidirectional gasification step, and on reversible
steam sorption as OH/H groups and irreversible surface chemistry [102]. A set of
kinetic rate laws of the Langmuir-Hinshelwood type are formulated to describe the
formation and consumption of each gas species [102]:

−K2 · pH2 O − K2 K3 · pH2 O pCO


rH2 O = (5.18)
1 + K4 pCO2 + K5 pH2 O
5.3. Chemical Kinetics 65

rH2 = −rH2O (5.19)


2K1 · pCO2 + K2 · pH2 O − K2 K3 · pH2 O pCO
rCO = (5.20)
1 + K4 pCO2 + K5 pH2 O
−K1 · pCO2 + K2 K3 · pH2 O pCO
rCO2 = (5.21)
1 + K4 pCO2 + K5 pH2 O
−K1 · pCO2 − K2 · pH2 O
rC = − (rCO + rCO2 ) = (5.22)
1 + K4 pCO2 + K5 pH2 O

ri is the reaction rate of species i ( i = H2 , H2 O, CO, CO2 and C) for heterogeneous


surface reactions defined as:
1 dni
− ri = (5.23)
mC dt
where mC is the mass of coke. In order to account for mass transfer limitations
inside the solid particle for varying particle diameter, the particle effectiveness η,
experimentally determined by [100], is introduced,

4 ·D 2 ·D
η (D) = 0.571e−1.29·10 + 0.429e−9.56·10 (5.24)

η converges to 1 for D approaching 0 and diffusion becoming instantaneous. The


temperature dependence of each Ki is determined by imposing the Arrhenius law
Ki (T ) = k0 exp −E

RT
A
. Apparent activation energies and frequency factors, cal-
culated by linear regression of experimental data obtained by thermogravimetric
measurements, are listed in Table 5.2 [100]. Comparison of the gasification rates be-
tween the experimental runs performed in the thermogravimeter (packed bed with
heating rates of 0.1–0.3 K/s by predominantly convective/conductive heat transfer)
and the experimental runs performed in the solar reactor (gas-particle entrained
flow with heating rates of 104 –106 K/s by predominantly radiative heat transfer, see
Chapter 7) led to the introduction of an empirical Arrhenius-type proportionality
constant ksolar = 498 · exp (−5.71 · 104 /RT ) to resolve the mass/heat transfer dif-
ferences between both set-ups and the fact that the release of volatiles by pyrolysis
occurs immediately at the entrance of the solar reactor. The volume-specific reaction
rate for species i and for a single particle diameter D can then be written as:

Ri = ksolar · η (D) · ri · M · YC · ρ (5.25)

For the polydisperse medium η is evaluated at the equivalent monodisperse diameter


66 5. Simulation Framework

Dη ,
R ∞
f (D) D3 · η (D) dD

−1 0
Dη = η R∞ (5.26)
0
f (D) D3 dD
where η −1 denotes the inverse function of eq. (5.24).
Chapter 6

System Analysis — Lumped


Parameters Model1

The engineering design of the reactor presented in the first part of this thesis was
supported by performance calculations based on a simple lumped-parameters model.
Figure 6.1 shows the mass and energy flows that are considered. The system bound-
ary is defined by the reactor shell. Coke and Argon are fed at ambient temperature,
whereas H2 O is first evaporated and fed at Tsteam =423 K. Inside the cavity the re-
actants are heated to the reaction temperature Tcavity and eventually react to form
a mixture of Ar, H2 , CO, CO2 and residual steam and carbon. The products exit
the reactor at Tcavity and are cooled down to Thx =473 K in a heat exchanger, where
a part of the sensible heat is recovered. A fraction of the incident solar power Q̇solar
is lost either by radiation, Q̇rerad , which accounts for reradiation from the cavity as
well as for emission and reflection from the window, or by conduction through the
cavity walls Q̇cond .

6.1 Governing Equations


A black box system delimited by the inner cavity walls is considered. The temper-
ature at any location inside the cavity and on the cavity walls is equal to Tcavity .

1
Material from this section has been published in: A. Z’Graggen, P. Haueter, D. Trommer, M.
Romero, J. C. de Jesus, and A. Steinfeld. Hydrogen production by steam-gasification of petroleum
coke using concentrated solar power — II. reactor design, testing, and modeling. International
Journal of Hydrogen Energy, 31:797–811, 2006.
68 6. System Analysis — Lumped Parameters Model

insulation reactor
oil cooling, Toil Tshell
window
Q̇chem
Q̇cond
Q̇rerad

Q̇solar Q̇heating Tcavity

coke

Q̇sensible
ṁcoke
ṅAr Thx
ṅAr
ṅH2 O Tsteam Ar
(1-XC) ṅC
Ar C
(1-XH2 O) ṅH2 O
H2O
Q̇steam XC ṅC
H2O T1 CO,H2

Figure 6.1: Diagram of the considered process and it’s energy and mass flows used
in the simulation.

Formulation of steady state energy conservation for the cavity results then in:
X X
0 = Q̇solar − Q̇rerad − Q̇cond + ṅi,in · hi (Tin ) − ṅi,out · hi (Tcavity ) (6.1)
species species

The enthalpy difference described by the last two terms of eq. (6.1) can also be
expressed as the sum of the power required to heat the feedstock from Tin to Tcavity ,
Q̇heating , and the enthalpy change due to the chemical reaction, Q̇chem :

0 = Q̇solar − Q̇rerad − Q̇cond − Q̇heating − Q̇chem (6.2)

where the conduction heat losses are calculated with the overall heat transfer coef-
ficient U presented in 5.1.2:

Q̇cond = Scone · Ucone · (Tcavity − Toil ) + Scavity · Ucavity · (Tcavity − T∞ ) (6.3)


6.1. Governing Equations 69

and Ucone = 65 and Ucavity = 22 W/(m2 K) are overall heat transfer coefficients
relative to Toil and T∞ , respectively. Values given are averages over Scone and Scavity ,
the inner surface of the reactor’s front cone and the surface of the rest of the cavity,
respectively.

The power lost by radiation Q̇rerad is given by:

Q̇rerad = Q̇solar · Rw (Tsun )


4
+ Saperture · σTcavity · Tr w (Tcavity )
4
Q̇solar · Aw (Tsun ) + Saperture · σTcavity · Aw (Tcavity )
+ (6.4)
2

where the first therm accounts for reflected incoming radiation, the second term
for radiation reradiated from the cavity and the third term for emission from the
window. The total window absorptance Aw , reflectance Rw and transmittance Tr w
are calculated from the spectral values presented in eqs (5.3)–(5.5) as:
R∞
0
Eλb (T ) · {Aλ,w , Rλ,w , Tr λ,w } dλ
{Aw , Rw , Tr w } = R∞ (6.5)
0
Eλb (T ) dλ

where Eλb (T ) is the Plank’s blackbody spectral emissive power at Tsun = 5780 K or
Tcavity depending on the origin of the radiation. The factor 1/2 in the third therm of
eq. (6.4) takes into account emission from both sides of the window. Note also that
the radiative flux incident on the window from the inside of the cavity, used in the
4
second and third term of eq. (6.4), is approximated simply by σTcavity . In fact the
apparent absorptions of the cavity is close to one and furthermore the radiation is
mostly absorbed/emitted by the particles. This is in contrast the method presented
previously in [123] where the radiosity method was used to determine incident fluxes
on the window.

The heat absorbed by the chemical reaction Q̇chem reads:

Q̇chem = XC · ṅC · ∆HR (Tcavity ) (6.6)

where the enthalpy of the reaction ∆HR is evaluated at the temperature of the
cavity. The gas composition at the outlet and the carbon conversion are found by
numerical integration of the system of differential equations given in eqs (5.18)–
(5.22), assuming a plug flow reactor. Similarly to Section 3.4.4 the fluid velocity is
70 6. System Analysis — Lumped Parameters Model

P 
given by u (t) = ṅ
species i (t) RT / (pS) and the integration is stopped at t = τ ,

where the mean residence time τ is given by the implicit relation L = 0 u (t) dt,
and L is the length of the reactor. The sensible heat that can be recovered from the
product gases is then defined as:

X Z Tcavity
Q̇sensible = ṅi · cp,i (T ) dT (6.7)
species Thx

and the heat used to produce the steam reads:


 Z Tsteam 
Q̇steam = ṅH2 O · (373.15 K − T∞ ) cp,H2 O(l) + ∆Hvap + cp,H2 O(g) (T ) dT
373.15 K
(6.8)
where ∆Hvap is the evaporation enthalpy for water.
Finally, eq. (6.2), which is implicit in Tcavity , is iteratively solved using the Nelder-
Mead Simplex method [69]. The temperature of the outer surface of the Inconel shell
Tshell , is calculated as:

Ucavity
Tshell = Tcavity − · (Tcavity − T∞ ) (6.9)
Ushell

where the overall heat transfer coefficient between cavity wall and outer Inconel shell
Ushell = 108 W/(m2 K) is calculated assuming 1D cylindrical conduction:

1
Ushell = P log(Di+1 /Di )
(6.10)
Dcavity · layers ki

where ki and Di describe the thermal conductivity and the diameter of the different
material layers involved, respectively.

6.2 Results
Two parameter studies were performed for the baseline simulation parameters listed
in Table 6.1: (1) the optimal volume and coke feeding rates were identified for a
reactor operating at 5 kW and with a 50 mm in diameter aperture, (2) the potential
of up-scaling the ideal configuration found in (1) was investigated up to a nominal
power of 1 MW.
Figure 6.2 shows the results for parameter study (1): the carbon conversion
6.2. Results 71

Table 6.1: Baseline model parameters for the studies conducted for the 5 kW lab
scale reactor and it’s scale-up.

Parameter study: lab scale scale-up


Aperture diameter Daperture , m 0.05 0.08–0.40
Cavity volume Vcavity , dm3 0.002–2.2 5.7–794
Feeding rates ṁcoke , g/min 0.6–20 31–1986
ṁH2 O , g/min 1.3–53 82–5257
V̇Ar , ln /min 2–66 0.0
Overall stoichiometric ratio H2 O/C, – 2.0 2.0
Solar power input Q̇solar , kW 5 40–1000
Static pressure p0 , Pa 105
Inlet equivalent particle diameter D30,in , µm 2.21
D32,in , µm 3.13
Overall heat transfer coefficient U , W/(m2 K) 22/65 10
Inlet temperature Tin , K 423.15
Temperature after the HX Thx , K 473.15
Oil temperature Toil , K 393.15 –
Surroundings temperature T∞ , K 293.15

at the exit of the reactor (a), defined in eq. (3.5), the solar-to-chemical efficiency
(c), defined in eq. (3.7), and the temperatures of the cavity (b) and of the Inconel
shell (d) as a function of the coke feeding rate and the reactor volume. A lower
petcoke feeding rate leads to a higher chemical conversion due to longer residence
times, but at the expense of a lower net solar energy absorbed. In contrast, a higher
coke feeding rates lead to a larger portion of energy used for heating the feedstock
but not converted into chemical energy. A similar effect has an increase in the
reactor volume, resulting in longer residence times and higher chemical conversion,
but at the expense of higher conduction losses. Optimal parameters determined for
maximum energy conversion efficiency were ṁcoke = 7.5 g/min and Vreactor = 0.54
dm3 . Unfortunately, these parameters resulted in very high temperatures for the
reactor liner Tcavity =1771 K as well as for the Inconel shell Tshell =1418 K. In order
to reduce the thermal load of the reactor components, the prototype reactor was,
therefore, designed with a cavity volume of 1.6 dm3 resulting in a slight decrease in
maximal ηchem from 20% to 17%.
The effect of scaling-up the reactor is elucidated in Fig. 6.3, where ηchem and
72 6. System Analysis — Lumped Parameters Model

(a) (b)
16 0.1
XC 1400

12 0.25
ṁcoke , g/min

1500
0.5
8 Tcavity, K
1600
0.9
2000 1800
4
1
2400 2200
0
(c) (d)
16 ηchem
0.05 0.05
0.1 Tshell, K
ṁcoke , g/min

12 1200
0.14
8 0.17 1300
0.19
1400
0.14 1600
4 0.1
0.05 1800
0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Vcavity, dm3 Vcavity, dm3

Figure 6.2: Carbon conversion (a), cavity (b) and shell (d) temperatures and chem-
ical efficiency (c) as a function of the carbon feeding rate and the reactor volume for
the 5 kW prototype reactor and the baseline parameter listed in Table 6.1.

ηproc together with the related Tcavity and XC are plotted as a function of Q̇solar and
ṁcoke , using as baseline the optimum operating conditions found in Fig. 6.2. The
aperture diameter and the volume were adjusted to keep a constant Q̇solar /Aaperture
2/3
ratio (average flux of 2300 suns) and constant Q̇solar /Vcavity , respectively. Further,
the oil cooling of the reactor front is omitted and an improved ceramic insulation is
assumed (U = 10, W/ (m2 K)). There is a remarkable positive effect of scaling-up the
reactor as a result of the relatively lower conduction losses through a smaller area-
to-volume ratio. For example, for an optimum petcoke feeding rate, the predicted
ηchem of a 40, 200, and 1000 kW reactor are 12%, 32%, and 35%, respectively,
whereas the predicted ηproc are 49%, 73%, and 74%, respectively. The locations of
maximal efficiency lie on straight lines of constant ṁcoke /Q̇solar in the range 0.13–0.15
6.2. Results 73

(a) (b)
120

1200
100

0.1

0.5

1000
9

0
0.
0.001

00
ṁcoke , kg/h

00

140
80 1

16
18
00
60 20
0
40 220
XC 2400
20
Tcavity, K
0
(c) (d)
120
ηproc

73
34

100

0.
0.
0.2

0.7
0.6
0.7
ṁcoke , kg/h

0.3 3

0.4
80 0.7
0.2
0.1

0.3

4
0.3
0.01

60
0.2 0.6
40
0.1 0.4
20 ηchem 0.2
0
200 400 600 800 1000 200 400 600 800 1000
Q̇solar , kW Q̇solar , kW

Figure 6.3: Predicted carbon conversion (a), cavity temperature (b), chemical effi-
ciency (c) and process efficiency (d) as a function of the carbon feeding rate and the
solar power for a scaled-up reactor. The baseline parameter are listed in Table 6.1.

kg/ (hr · kW) with typical related XC in the range 0.9–1.0 and Tcavity in the range
1700–1900 K.
Finally, the simulation model was run for the operational parameters of each
experimental run listed in Tables 4.1 and 4.2 for Campaign 1 and Campaign 2, re-
spectively. Figure 6.4 shows numerically calculated and experimental measured data
for the (a) 23 solar experimental runs of Campaign 1; and (b) 20 solar experimental
runs of Campaign 2, ordered by increasing Q̇solar . Shown are the cavity and shell
temperatures, the carbon conversion and the steam conversion. Full circles indicate
numerically calculated values; open circles indicate the experimentally measured
data. The error bars indicate the propagated inaccuracy of the input parameters,
namely ±913 % and ±5 % for the solar input power Q̇solar and the feeding rate ṁcoke ,
74 6. System Analysis — Lumped Parameters Model

(a) temperature, input power (b) temperature, input power

T cavity T cavity 4.5


6 1600
1600
5.5 4.4

Q̇ solar, kW

Q̇ solar, kW
5
T, K

T, K
1200 4.3
1200
4.5
T shell T shell 4.2
4
800 RMST 800 RMSTcavity = 0.018
cavity = 0.016
RMSTshell = 0.05 RMSTshell = 0.045

carbon conversion carbon conversion


1 1

0.8 0.8

0.6 0.6
XC

XC

0.4 0.4

0.2 0.2
RMS = 0.31 RMS = 0.63
0 0
steam conversion steam conversion
1 1
RMS = 1.3 measured
0.8 0.8 calculated

0.6 0.6
X H2 O

X H2 O

0.4 0.4

0.2 0.2
RMS = 0.47
0 0
1 5 9 13 17 21 1 5 9 13 17
experiment # experiment #

Figure 6.4: Numerically calculated (lumped parameters model) and experimental


measured data for the (a) 23 solar experimental runs of Campaign 1; and (b) 20
solar experimental runs of Campaign 2, ordered by increasing Q̇solar . Shown are the
average cavity and wall temperatures, the carbon conversion and, the chemical and
process efficiencies. Full circles indicate numerically calculated values; open circles
indicate the experimentally measured data; the error bars indicate the propagated
inaccuracy of the input parameters.

respectively (see Section 3.3). The simulation was able to successfully predict the
reactor temperatures. In fact, relative RMS errors below 5% were found for the
shell temperatures Tshell and the cavity temperatures Tcavity and for both campaigns.
6.2. Results 75

In Campaign 1, most predictions for XC and XH2 O lie inside the accuracy bounds.
Nevertheless, the considerable width of these accuracy bounds shows the fuzziness
of the results. For Campaign 2, finally, the simulation fails to predict the chemical
conversion, especially for H2 O, the RMS being 130%. In fact, the simple model used
does not consider the dependency between particle size and radiation absorption ef-
ficiency, thus overestimating the performance of experimental runs performed with
poorly absorbing particles (experimental Campaign 2 with feedstock type 3).
76 6. System Analysis — Lumped Parameters Model
Chapter 7

Heat and Mass Transfer in the


Reactor Cavity1

A model for heat and mass transfer inside the cavity of the reactor presented in
Section 3.1 was developed. It solves for radiative, conductive, and convective heat
transfer, fluid flow and reaction kinetics. The model is implemented as two separate
solver modules: (1) a finite volume solver (CFD) that solves for coupled radiative,
conductive and convective heat transfer, fluid flow and reaction kinetics and (2) a
Monte-Carlo (MC) raytracer that provides the source term for radiation used in (1).
The intrinsic statistical variance induced by the MC solver is reduced by means of
Gaussian kernel smoothing and adaptive underrelaxation when coupled to the finite
volume solver. The main boundary conditions were previously presented in Chapter
5 as well as the treatment of the walls and the window.
The two-phase medium considered is composed by a solid particulate polydis-
persion suspended in a gas mixture. The particle diameters of the solid phase are in
the range 1–100 µm. The particle suspension is modeled as a non gray absorbing,
emitting, and scattering participating medium subjected to concentrated thermal
radiation. Typical radiative equilibrium temperatures are in the range 1500–1800
K. Both solid and gas phases are involved in the heterogeneous chemical reaction
taking place predominantly on the outer surface of the solid particles and, to some
extent, on the inner porous surface of the particles. As the gasification reaction
1
Material from this chapter has been submitted for publication as: A. Z’Graggen, and A.
Steinfeld. Heat and Mass Transfer Analysis of a Suspension of Reacting Particles subjected to
Concentrated Solar Radiation — Application to the Steam-Gasification of Carbonaceous Materials.
International Journal of Heat & Mass Transfer, 2007.
78 7. Heat and Mass Transfer in the Reactor Cavity

progresses, the solid particles shrink and their thermal properties vary, as well as
the optical properties of the particle suspension.
The scheme developed here is able to deal with concentrated thermal radiation
input, polydisperse particulate media with spectral and directional optical proper-
ties, and temperature-dependent chemical kinetics and fluid properties. Validation
is accomplished by comparing numerical and experimental results.

7.1 Heat Transfer Modes in the Polydisperse Par-


ticle Suspension
Typical relaxation times are compared for assessing the relative importance of the
different modes of heat transfer, namely conduction, convection, and radiation. The
reference time scale for the fluid flow, τflow = L/u , equals 0.1 s. For conductive heat
transfer inside the particles, the relaxation time of the core temperature of a sphere
with respect to a sudden change in its surface temperature is given by [44]:

ρs · cp,s · D2
τcond = (7.1)
4 · ks

Values 10 to 105 times smaller than τflow are found for particles of 100 to 1 µm,
respectively. Uniform temperature is therefore assumed for a single particle. The
relaxation time for convective heat transfer of the temperature of a sphere submerged
in a fluid at a different temperature reads:

ρs · cp,s · D
τconv = (7.2)
6·h

Particles are assumed to be entrained by the gas flow, as justified by Stokes numbers
Stk< 10−2 , thus the Nusselt number for a sphere is Nu=2 [44] and eq. (7.2) simplifies
to:
ρs · cp,s · D2
τconv = (7.3)
12 · kg
The resulting τconv is comparable to τflow for 100 µm particles. The energy conserva-
tion equation is therefore solved for each phase separately. Finally, the temperature
T of a particle homogenously irradiated is described as a function of time by:

πD3 ρs cp,s ∂T
= πD2 σ T∞
4
− T4

· (7.4)
6 ∂t
7.2. Governing Equations 79

where T∞ is the temperature of the surroundings. Linearization of T 4 around a


typical particle temperature T0 leads to:

T 4 = T04 + 4T03 T (7.5)

The relaxation time of the temperature of a particle homogenously irradiated is


then:
ρs · cp,s · D
τrad = (7.6)
24 ·  · σ · T03
Calculated τrad are 1 to 3 orders of magnitude smaller than τflow . The temperature
field is therefore mainly influenced by the local divergence of the radiative flux and
only marginally by the heat transported by advection.

7.2 Governing Equations


The solid particles are assumed to be entrained in the gas phase, as justified by
the Stokes number in the range of 10−2 –10−6 . Conduction and diffusion in the flow
direction are neglected, as justified by the Peclet numbers for conduction (Pe =
uL/α) and for diffusion (Pe = uL/Dc ) in the order of 102 . Steady-state mass
conservation for the mixture is then expressed by:

∇ · (ρ · u) = 0 (7.7)

and for a single species i by:


∇ · (ρYi · u) = Ri (7.8)

where Ri denotes the volumetric rate of production/consumption of species i as


defined in Section 5.3. Steady-state energy conservation for the solid phase yields:

∇ · (ρs hs · u) = φrad − φconv − φchem (7.9)

where the radiation source φrad is given by the negative divergence of the radiative
flux −∇ · qr , the convection source φconv is given by the convective heat exchange
between the two phases, and the chemistry source φchem account for enthalpy carried
by mass changing its phase. Steady-state energy conservation for the gas phase
yields,
∇ · (ρg hg · u) = φconv + φchem (7.10)
80 7. Heat and Mass Transfer in the Reactor Cavity

The mean absorption coefficient of a typical gas composition (CO, CO2 , H2 , H2 O)


at 1 bar, calculated line-by-line using the HITEMP/HITRAN database [83], is ap-
prox. 0.06 and 0.006 m−1 for radiation emitted at 1800 and 5780 K, respectively.
Absorption by the gas is therefore neglected vis-à-vis absorption by the particles
(see Appendix C for details).

Energy source for convection Convective heat transfer between the polydis-
perse solid and gas phases is described by Newton’s law of cooling [44] as the sum
of each particle’s contribution:

h · πD2
P
Np
φconv = (Ts − Tg ) (7.11)
V
Np π R∞
further using h = Nu · kg /D and fV = 6V 0
f D3 dD leads to:
R∞
6 · fV · Nu · kg 0 f D dD
φconv = R∞ (Ts − Tg ) (7.12)
0
f D3 dD

Finally, the Nusselt number for a stationary sphere is equal to 2 [44] and the defini-
tion of equivalent diameters (eq. (5.10)) is used:

12 · fV · kg
φconv = 2
· (Ts − Tg ) (7.13)
D31

Energy source for chemistry The mass consumed, generated, or transformed


during the chemical reaction contributes to the enthalpy change, evaluated at the
temperature of the originating phase:
n
X
φchem = hi (T ) · Ri (Ts , Yi , f ) (7.14)
i=1

Note that the rate of reaction Ri depends on the temperature of the solid phase,
Ts , the gas composition, and the particle size distribution f . At the high tem-
peratures considered, only two reactions are of importance, namely gasification
C + H2 O → H2 + CO and water-gas-shift CO + H2 O → H2 + CO2 . Equation (7.14)
can then be explicitly written as:
7.2. Governing Equations 81

φchem = hH2 O (Tg ) · RH2 O (7.15)


+hH2 (Ts ) · RH2
+hCO (Ts ) · (RCO − RCO2 ) + hCO (Tg ) · RCO2
+hCO2 (Tg ) · RCO2

where the reaction rates Ri are calculated with equations (5.18)–(5.22). Note that
the temperature field is related to the enthalpy by the implicit equation:
Z T
h (T ) = href + cp (T ) dT (7.16)
Tref

where href is the reference enthalpy at Tref = 273 K, and temperature-dependent


values of cp for all species are calculated with the empirical correlations given in
[88].

Energy source for radiation The divergence of the radiative flux is obtained
from the difference between emitted and absorbed radiation
Z ∞  Z 4π 
1
∇ · qr (s) = 4π κλ (s) Iλb (s) − Iλ (s, ω) dω dλ (7.17)
λ=0 4π ω=0

Using the Planck mean absorption coefficient κP eq. (7.17) can be simplified to:
Z ∞ Z 4π
4
∇ · qr (s) = 4κP σT − κλ (s) Iλ (s, ω) dωdλ (7.18)
λ=0 ω=0

where the first term accounts for emission from the volume and the second term for
absorption in the volume. The latter is further referred to as φrad,a . The radiative
intensity Iλ , required in the second term, is given by the equation for radiative
transfer for a participating medium
Z 4π
dIλ σsλ (s)
= −βλ (s) Iλ (s) + κλ (s)Iλb + Iλ (s, ωic ) Φλ (ω, ωic ) dωic (7.19)
ds 4π ωic =0

solved by Monte-Carlo raytracing (see Section 7.3.1). Note that the solution of
eq. (7.19) in a bounded simulation domain also indirectly yields values for the radia-
R∞ R∞
tive power absorbed on the walls 0 q̇ic,λ λ dλ and by the window 0 q̇ic,λ Aw,λ dλ,
as introduced in eqs (5.1) and (5.6), respectively.
82 7. Heat and Mass Transfer in the Reactor Cavity

7.3 Numerical Implementation

selective quartz window: solid and gaseous product


- absorbing, emitting and feedstock at Tin outlet
transmitting radiation - initial composition
- conduction losses - initial velocity cylindrical
compartments
5
4
two-phase medium: 3 r
- polydisperse 2
1 x
- chemically reacting
- absorbing, emitting
and scattering

concentrated solar power: ceramic insulation:


- solar spectrum - absorbing, emitting and
- angular distribution reflecting radiation (diffuse-gray)
- radial distribution - conduction losses

Figure 7.1: Axis-symmetric model domain, featuring five concentric cylindrical com-
partments. Indicated are also the boundary and inlet conditions. The grid spacing in
x-direction is adapted to the expected temperature gradients, shown is an example
for the Monte-Carlo solver.

The simulation domain and the main boundary conditions are shown in Fig. 7.1.
Preliminary CFD simulations showed an axis-symmetric flow pattern with negligible
radial velocity components. However, in order to account for the radial dependency
of the incoming solar radiation (see Fig. 5.2), the domain is subdivided into con-
centric cylindrical compartments, as indicated in Figs 7.1 and 7.2, where a typical
finite volume is described. As shown is Section 7.1 advective heat transport is of sec-
ondary importance. Thus, a two-phase fluid flow with neglected angular and radial
components of the velocity vector is assumed. This simplification further enables
to better elucidate the physical phenomena involved in the interaction of radiation
with the chemical reacting flow. The homogenous grid spacing in the radial di-
rection (number of compartments) is equal for both, the MC and the CFD solver.
In contrast, the grid spacing in the x-direction is adapted to the requirements of
each solver separately. Each volume element is considered isothermal (each phase
separately), with homogenously distributed species and particle size fractions. The
following boundary/initial conditions and material properties are set:
7.3. Numerical Implementation 83

advection
compartment j+1 Sj+1 radiation

compartment j Sj P D, T, fV

u, ½, Yi, h
w e

compartment j{1 Sj{1


r
¢x x

Figure 7.2: Typical finite volume considered in the simulation with faces ‘w’ and ‘e’,
centerpoint ‘P’, finite length ∆x, and cross-section Sj . Also shown are the respective
face-centered and cell-centered variables. Advection is solved in x-direction for each
compartment j separately, whereas radiative exchange is considered along both, the
x and the r directions.

• Concentrated solar radiation The angular and radial distribution of the in-
coming radiation were presented in Section 5.1.1. The spectral distribution is
approximated by Planck’s blackbody emission at 5780 K.

• Quartz window Spectral values for the transmittance, absorptance and re-
flectance of the window, as well as the overall heat transfer coefficient for
conduction/convection losses were given in Section 5.1.3. In contrast to the
experimental reactor, where the window is mounted on a diverging frustum
away from the aperture, in the model the window is located at the aperture
for two reasons: (1) the radiation BC is given at the aperture and (2) the com-
putational domain is reduced. This simplification does not affect the results
since no radiation-particle interactions occur in the diverging frustum.

• Reactor walls Overall heat transfer coefficients for conduction in the walls as
well as radiative properties were given in section 5.1.2. The walls are considered
diffuse gray ( = 0.8) due to particle deposition (visual observations). The
impact of the uncertainty of  on the results was assessed by sensitivity analysis
84 7. Heat and Mass Transfer in the Reactor Cavity

(see Section 7.4).

• Feedstock injection The inlet axial velocity is assumed to be uniformly dis-


tributed over the cross section, as justified by preliminary CFD simulations
and visual observation. Further assuming ideal gas, the initial velocity is given
by:
(ṅH2 O + ṅAr + ṅH2 ,feedstock ) · R · Tin
uin = (7.20)
p0 · S
where p0 is the ambient pressure, S is the cavity cross-sectional area and
ṅH2 ,feedstock is the hydrogen bound in the feedstock released by fast pyrolysis
at the entrance of the reactor.

• Solid particles Petrozuata delayed coke was used as feedstock (Chapter 2).
Particles size distribution and radiative properties were given in Section 5.2.
Equivalent monodisperse diameters were used to describe the changing prop-
erties of the polydisperse medium as a consequence of the chemical reaction.
Reaction kinetics were presented in Section 5.3.

7.3.1 Monte-Carlo Ray Tracer


An overview of common approaches used to deal with radiative heat transfer in
participating media is given in [43] and [96]. The Monte-Carlo (MC) method used
in this study was chosen for it’s simplicity and flexibility. It permits to virtually
model any complex physical phenomenon that affects radiative transfer, including
non-isothermal, non-gray and anisotropically-scattering media. Furthermore, com-
plex geometries, coupled convection and conduction heat transfer as well as spa-
tially varying medium properties can be easily implemented. Previous pertinent
MC modeling studies include pulverized coal furnaces [33] and solar reactors for
coal gasification [60, 110] and CH4 decomposition [40].
The source term for radiation absorbed in the volume, φrad,a (second term of
eq. (7.18) describing the divergence of the radiative flux), as well as the radiation
absorbed by the walls q̇ic is calculated with a statistical simulation in which the his-
tory of a large number of rays either emitted in the volume and on the boundaries
or incident through the window is tracked. Emission wavelength and direction, as
well as absorption and scattering of these rays is determined by means of probabil-
ity density functions (PDF) based on local radiative properties and using random
7.3. Numerical Implementation 85

select ray starting point


new ray: qray, λ
select ray direction: r
compute penetration
length: l(βλ,s) or ln(βλ,s)
compute distance to next
face: Lf
ray starting
update point
hit =
dimensionless inside element? location of
penetration l<Lf particle
length no yes interaction

ray starting compute absorption compute absorption


point by the gas phase: by the gas phase:
= Ea,g(κλ,g,l) Ea,g(κλ,g,l)
intersection
with face update qray update qray

qray>qmin qray>qmin
no
yes yes

compute compute
reflection face is scattering
direction boundary? direction rs

compute absorption
ray starting yes by the solid phase:
point Ea,s(ωλ)
= compute absoption
intersection at the boundary update qray
with
boundary update qray
qray>qmin yes
yes qray>qmin no

Figure 7.3: Flowchart of the Monte-Carlo raytracer used the calculate the radiative
source term in a domain composed of finite volumes. Note that the volumes are
delimited by either interfacial faces or/and domain boundaries. Computation steps
that involve one or several random numbers are show by the bold boxes.
86 7. Heat and Mass Transfer in the Reactor Cavity

numbers R drawn from a uniformly distributed set between 0 and 1. The sum of
all ray’s interactions with the two phase medium and the walls yields then a spatial
distribution of absorbed radiation:
P
rays Ea,wall
q̇ic = (7.21)
P S
rays Ea,volume
φrad,a = (7.22)
V

where Ea,wall and Ea,volume describe each ray’s contribution to the power absorbed
on a wall element dS and in a volume element dV , respectively. The simulation
domain shown in Fig. 7.1 is dived in hexahedral elements; due to the rotational
symmetry of both the flow field and the radiative BC only a slice of the reactor
is considered, while the cutting planes are modeled as perfect reflecting mirrors.
Figure 7.3 shows the flowchart of the Monte-Carlo approach used [113, 24, 119].
Rays are generated by emission from the walls and the volumes (gas and solid phase)
as well as by transmission through the window (incident q̇solar ). Each ray is described
by its power qray , wavelength λ, starting point and direction r. A ray traveling in
a volume dV can undergo four different interactions: (1) continuous absorption in
the gas phase, (2) local absorption and scattering in the solid phase, (3) absorption,
reflection and transmission on a wall element dS and (4) transmission to the adjacent
volume element. For each absorption interaction the ray’s power is diminished by
the amount Ea related to local radiative properties. The ray history is terminated
0
P
when the residual power qray = qray − Ea is smaller than the cutoff power qmin or
when it leaves the reactor through the window. Note that the treatment of radiation
absorption in the gas phase is presented for completeness, although it is neglected
in the simulation results presented later on. The equations describing the behavior
of the above mentioned interactions are divided into three groups.

Interaction with a wall The total power of the rays emitted from a wall element
is given by:
4
E = S ·  (T ) σTwall (7.23)

where  is the total emissivity of the surface. The ray starting points are homoge-
nously distributed on the surface element dS while the direction of the emitted rays
is given by the azimuthal angle θ and the circumferential angle φ related to the face
7.3. Numerical Implementation 87

normal n [91] (see also Fig. 7.4):

φ = 2 · π · Rφ (7.24)
Rθ R∞ 0 0 0

2π 0 0
 λ (λ, θ ) Iλb (λ, Twall ) sin θ cos θ dλ dθ0
Rθ = 4
(7.25)
σTwall

where λ (λ, θ) is the spectral directional emissivity of the surface. The wavelength
λ of the emitted ray is given by [91]:
Rλ R π/2 
2π 0
Iλb (λ0 , Twall ) 0
λ (λ0 , θ) sin θ cos θ dθ dλ0
Rλ = 4
(7.26)
σTwall

The power absorbed by the wall when hit by a ray is given by Ea = qray · λ (λ, θ).
The direction of the reflected ray, which carries the remaining power qray − Ea , is
either calculated with eqs (7.24) and (7.25) for a diffuse surface or by rreflected =
r − 2 (r · n) n for a perfectly specular surface while λ does not change.

Interaction with a window The window is treated similarly to the walls. The
spectral directional emissivity λ (λ, θ) is simply substituted for the spectral direc-
tional absorptance Aw (λ, θ) of the window when calculating emission and absorp-
tion. The amount of power reflected is calculated with the spectral directional
reflectance of the window Rw (λ, θ). In addition, radiation incident from the outside
of the domain is considered. The radial, directional and spectral distribution of the
incident solar power is given in Section 5.1.1. A part of the solar power is absorbed
and reflected (q̇solar · Aw (Tsun ) and q̇solar · Rw (Tsun )). The remaining power enters
the cavity while the angular distribution shown in Fig. 5.3 is weighted by Tr w (λ, θ)
calculated with eq. 5.3.

Interaction with a volume The gas and particulate phases are assumed perfectly
mixed and homogeneous over the whole volume. The total emitted power is given
by:
E = Eg + Es = 4V · κP,g (Tg ) · σTg4 + κP,s (Ts ) · σTs4

(7.27)
R∞
where κP is the Planck mean absorption coefficient defined as 0 κλ Eλb dλ/σT 4 and
the absorption coefficient κλ of the solid phase is given in Section 5.2. The emission
88 7. Heat and Mass Transfer in the Reactor Cavity

volume element: intersection with the


·¸,g boundary if no
¯¸,s=·¸,s+¾s¸,s scattering had
occured
Lf r
boundary element:
µ transmission to
"¸(µ), Tw¸(µ) adjacent element
l rs
Á
emission from absorption and
the wall or Ár
µ scattering by the
transmission of solid phase
incident solar n
radiation: absorption by
qray, ¸ the gas phase

Figure 7.4: Typical volume element used in the MC solver. A ray emitted from a
boundary face (gray) with initial power qray and wavelength λ undergoing absorption
and scattering is shown (dotted arrow). The participating medium is described by
κλ,g , κλ,s , and σsλ,g the absorption coefficients for the gas and solid phase, and the
scattering coefficient of the solid phase.

direction is given by

φ = 2 · π · Rφ (7.28)
θ = cos−1 (1 − 2 · Rθ ) (7.29)

while the wavelength is calculated as:


Z λ
1
Rλ = κλ Eλb dλ0 (7.30)
κP · σT 4 0

Figure 7.4 shows a typical volume element considered in the MC simulation. A


ray traveling trough the volume is either absorbed by the gas phase or absorbed
and scattered by the solid phase. Absorption by the gas is assumed significantly
smaller than by the particles. The penetration length of a ray, i.e., the distance
until extinction by the solid phase occurs, l, is found from:
Z l
− βλ,s (s) ds = ln Rl (7.31)
0
7.3. Numerical Implementation 89

ray-particle
interaction
ray
ln

¯1 ¯2 ¯n
L1 L2 ... Ln

Figure 7.5: Schematic of a ray passing trough several elements of width Ln and
extinction coefficient βn . Also shown is the residual penetration length ln in the last
element.

where the extinction coefficient βλ,s = κλ,s +σsλ,s of the solid phase is given in Section
5.2. For a homogenous volume, eq. (7.31) simplifies to l = − ln (Rl ) /βλ,s . The power
absorbed by the gas phase along distance l is then given by:

Ea,g = qray (1 − e−κλ,g ·l ) (7.32)

while the power absorbed by the solid phase at a distance l from the emission point
is expressed as [113]:
Ea,s = qray (1 − ωλ ) (7.33)

where the scattering albedo ωλ is defined as ωλ = σsλ,s / (κλ,s + σsλ,s ). Finally, the
direction of the ray after scattering is given by:

1 θ
Z
Rθ = Φ (θ0 ) sin θ0 dθ0 (7.34)
2 0
φ = 2πRφ (7.35)

where θ is the angle between the forward direction of the incident ray r and the
direction of the scattered ray rs , φ is given on a plane perpendicular to r, and the
phase function Φ (θ) of the polydisperse solid phase is given in Section 5.2.
In a medium with a small extinction coefficient or when very small volumes are
considered a ray is likely to pass trough multiple volume elements before experiencing
the first interaction with the solid phase (see Fig. 7.5). In this case eq. (7.31) is solved
for piecewise constant properties. Assuming absorption in the nth element it reads
90 7. Heat and Mass Transfer in the Reactor Cavity

then: Z l
L1 β1 + L2 β2 + · · · + Pn−1
βn (s) ds = − ln (Rl ) (7.36)
i=1 Li
Pn−1
The remaining path length in element n, ln = l − i=1 Li is then given by:

− ln (Rl ) − L1 β1 − L2 β2 − · · · − Ln−1 βn−1


ln = (7.37)
βn

where the numerator describes the dimensionless penetration length, − ln (Rl ), de-
creased by the sum of the optical thicknesses of the traversed layers. Note that a new
dimensionless penetration length is calculated each time that a ray hits a particle
or a wall but not when it hits a specular symmetry boundary.

7.3.2 Fluid Flow Solver

The fluid flow is solved in a array of concentric cylindrical compartments as shown


in Figs 7.1 (5 compartments) and 7.2. Furthermore, the angular component of the
flow is neglected, thus eqs (7.7)–(7.10) — solved for each concentric comportment j
separately — are reduced to the sole axial dimension x. The resulting flow field is
of parabolic type (one way coordinate) [74] and is therefore solved directly volume-
by-volume starting from the inlet. Equations (7.7)–(7.10) are discretised into finite
R H
volumes, taking advantage of the relation V ∇·F dV = δV F n dS. The 1D volumes
are described by a cross-section S and finite length in the flow direction ∆x and are
delimited by the face ‘w’ at the inlet, and the face ‘e’ at the outlet. Mass conservation
of the mixture results then in (index j omitted for clarity):

(ρu)w
ue = (7.38)
ρe

while mass conservation for each species i yields:


Z
1
(ρuYi )e − (ρuYi )w = · Ri dV = ∆x · Ri (7.39)
S V

A downwind scheme is used to estimate the average reaction rate Ri over the control
volume since the reaction rate has a negative exponential behavior. Best convergence
was found for Ri = 3/8 · Ri,w + 5/8 · Ri,e . The mass fraction of species i at face ‘e’
7.3. Numerical Implementation 91

is then:
(ρuYi )w + ∆x( 83 Ri,w + 58 Ri,e )
Yi,e = (7.40)
ρe u e
Energy conservation for the solid and the gas phase are treated similarly:

(ρs uhs )w + ∆x φrad − φconv − φchem
he,s = (7.41)
(ρs u)e

(ρg uhg )w + ∆x φconv + φchem
he,g = (7.42)
(ρg u)e

where the average source terms for radiation and convection are calculated by the
central scheme φrad = 0.5 (φrad,w + φrad,e ) and φconv = 0.5 (φconv,w + φconv,e ), while
the average source term for chemistry is calculated similarly to the average reaction
rate: φchem = 3/8 · φchem,w + 5/8 · φchem,e .
Equations (7.38)–(7.42) are solved iteratively for each finite volume consecutively.
The step length ∆x is adapted to ensure conversion after 8 iterations. Due to the
stiffness of the system of differential equations that describe the chemical reaction
(eqs (5.18)–(5.22)) typical finite volumes for the CFD solver are 10 times smaller
(∆x ≈ 0.1 mm) than those used by the Monte-Carlo raytracer. Furthermore, when
particles in the 1 µm range are considered, particle-gas convection becomes virtually
instantaneous (τconv  τflow ) and volumes that are up to 105 smaller than those used
by the Monte-Carlo raytracer are required to ensure convergence.

7.3.3 CFD-MC Coupling

Figure 7.6 shows the flowchart of the coupled simulation setup. The iterative process
is started with a guessed temperature and mass distribution inside the cavity. The
MC solver calculates then the source terms φrad,a (x) and q̇ic (x) required by the
CFD solver. The latter solves for heat and mass conservation returning values for
Tg (x), Ts (x) D32 (x) and fV (x) to the MC solver. This procedure is repeated until
the relative change in the results between two consecutive iterations is smaller than
10−5 .
Note that, because of the statistical noise, the time required to perform a single
iteration at a reasonable accuracy for the MC module is several orders of magnitude
larger than that for the CFD module. For example, a simulation with 10 com-
partments run on an Intel R Pentium R 4 CPU at 2.53 GHz requires 1.2 seconds
92 7. Heat and Mass Transfer in the Reactor Cavity

guess temperature and


mass distribution

solve radiative transfer


equation by MC raytracing

calculate divergence of calculate incident radiative


radiative flux in the volume power on the walls
iterate

apply smoothing filters and


adaptive underrelaxation

solve coupled energy and mass


conservation equations

re-calculate temperature and mass re-calculate


distributions, and equivalent temperature
diameters in the volume distribution of the walls

Figure 7.6: Flowchart of the coupled solver. The bold box describes the MC ray-
tracer described in Fig. 7.3.

for one iteration of the CFD module. The MC module of the same simulation —
solved on a grid composed by 1577 volume elements and 5663 faces — requires 4,
29, and 297 seconds for 104 , 105 , and 106 primary rays (each primary ray results
in approximately 150 follow-up rays until its power is completely absorbed) with a
related standard deviation for the radiative source term of 0.12, 0.04, and 0.014, re-
spectively. To some extent, this problem is alleviated by Gaussian kernel smoothing
with adaptive bandwidth. For the incident radiation on the walls, the local mean
estimator is expressed by [12]:
Pn
w (xi − x, hi ) · q̇ic,MC (xi )
q̇ic0 (x) = i=1P
n (7.43)
i=1 w (xi − x, hi ) ·

where q̇in,MC denotes theresult


.of the MC module, w is the Gaussian kernel de-
2 √
fined as w (x, hi ) = exp −x

2
2hi
h i 2π , and the bandwidth hi is proportional
to the local grid spacing. The second term of the divergence of the radiative flux,
φrad,a is estimated analogously to eq. (7.43), where a two-dimensional kernel of the
7.3. Numerical Implementation 93

0
10
1
–1
10

Pemit /max (Pemit )


0.8
RMSrel; γ
–2
10 0.6
–3
i0
10 0.4
–4
10 RMSrel 0.2
γ
–5
emitted power
10 0
0 100 200 300
i

Figure 7.7: Relative RMS error for the temperature and underrelaxation parameter γ
(left axis) and total emitted power, scaled by its maximal value, used as an indicator
for system convergence (right axis) as function of the iteration number i. Values for
run #12 of Campaign 1 with ξ = 32.

form w (xi − x, hx ) · w (ri − r, hr ) is used with bandwidths hx and hr proportional


to the local grid spacing in the x and r direction, respectively [84]. Gaussian kernel
smoothing is not intrinsically conservative, the resulting radiative source terms q̇ic0
and φ0rad,a are therefore corrected:
R

q̇ic = q̇ic0 · R ic,MC
walls
(7.44)
q̇ 0
walls ic
R
φrad,a,MC
φrad,a = φ0rad,a · volume
R (7.45)
φ0
volume rad,a

An adaptive underrelaxation scheme for the radiative source term is introduced to


ensure convergence and reduce the overall computational time:

φi = 1 − γ i φi−1 + γ i φiMC

(7.46)

where i is the iteration step, φiMC denotes the result of the MC module (either the
source term in the volume or on the walls), and γ is the adaptive underrelaxation
factor. Initially, γ is set to γ 0 = 1 and the radiative source term is calculated
94 7. Heat and Mass Transfer in the Reactor Cavity

with low accuracy but quickly. The total power emitted from the volumes and
P P
the walls by the MC solver Pemit = volumes E + walls E is used to monitor the
system convergence. Once Pemit as a function of the iteration number i approaches
a constant value, convergence is reached and γ is subsequently reduced according
to:  
i 0 i0 − i
γ = γ exp (7.47)
ξ
where i0 is the iteration step for the system convergence and ξ is a positive real
number. Figure 7.7 shows the development of the relative RMS error for the particle
temperature:
 !2 , 0.5
Nvolumes
X Tji+1 − Tji
RMSrel =  Nvolumes  (7.48)
j
Tji+1

A MC simulation with only 4 · 104 primary rays results in convergence after 150
iterations, yielding a relative standard deviation of 0.11 and 0.07 for the raw and
the smoothed φirad,a , respectively. Significant increase in accuracy is accomplished by
reduction of γ beyond iteration step i0 (ξ = 32) as observed in the reduction of the
RMSrel and in the flattening out of Pemit . Alternative variance reduction techniques
used in computer graphics, as for example Quasi Monte Carlo Sampling [67, 56]
and Tiled Blue Noise Samples [38], were applied but did not enhance the simulation
results for radiative transfer calculations.

Two distinct grid convergence studies were performed to asses the discretization
error. Operational parameters of a typical run from experimental Campaign 1 were
used. In the first study grids composed by 8 concentric compartments (radial direc-
tion) and 10, 20, 40, 80 and 160 elements along the axial direction were considered.
In the second study grids composed by 2, 4, 8, 16 and 32 concentric compartments
and 80 elements in the axial direction were considered. Figure 7.8 shows relative
errors for XC and XH2 O as a function of subsequent grid refinement in axial direction
x and in radial direction r defined as:

|Xi − Xi,finest |
= (7.49)
Xi,finest

were Xfinest describes the result for 160 axial elements and 32 radial compartment
for studies one and two, respectively. Discretization errors smaller than 10−3 are
7.4. Results and Validation 95

XC
XH2 O
{2 ²x
10 ²r

²
{3
10

{4
10

1 2 4 8
grid refinement factor

Figure 7.8: Relative grid convergence errors in radial direction, r , and in axial
direction, x , for carbon and steam conversions at the outlet as a function of the
grid refinement factor. The grid refinement factor 1, 2, 4 and 8 corresponds to 10,
20, 40, 80 elements in x-direction and to 2, 4, 8 and 16 elements in r-direction.

obtained for grids with more than 40 elements in axial and 10 compartments in
radial direction. A grid convergence index (CGI) of 0.07% (Fs = 3) and an order
of convergence p ≈ 2 is found for the axial direction [80]. In order to account
for differing operational conditions (in particular with respect to the absorption
coefficient) the minimal grid size in axial direction is doubled to 80 elements.

7.4 Results and Validation


Sets of 23 and 29 solar experimental runs were carried out for Campaigns 1 and 2,
respectively (Chapter 4). The 9 experiments performed with slurry made of feed-
stock types 2 and 4 for Campaign 2 are not considered here, because, due to the
occasional clogging of the feeding system, no reliable values for ṁcoke were obtained.
Numerically simulated and experimentally measured cavity and reactor shell tem-
peratures, and coke/steam conversions at the reactor outlet are shown in Figs 7.9
(a) and (b) for Campaign 1 and Campaign 2, respectively. Tcavity for the experi-
mental runs is the mean between the pyrometer reading corrected for the windows’
transmittance and the thermocouple reading Tshell corrected for conduction through
96 7. Heat and Mass Transfer in the Reactor Cavity

temperature, input power (b) temperature, input power


(a)
2000 4.5
T cavity T cavity
6 1800
1600 4.4

Q̇ solar, kW
Q̇ solar, kW
1600 5.5

T, K
T, K

1400
5
1200 1200 4.3
T shell 4.5 T shell
1000
4 4.2
800 RMSTshell = 0.056 RMSTshell = 0.013

carbon conversion carbon conversion


1 1

0.8 0.8

0.6 0.6
XC
XC

0.4 0.4

0.2 0.2
RMS = 0.2 RMS = 0.25
0 0
steam conversion steam conversion
1 1
RMS = 0.18 RMS = 0.39 measured
0.8 calculated
0.8

0.6 0.6
X H2 O
X H2 O

0.4 0.4

0.2 0.2

0 0
1 5 9 13 17 21 1 5 9 13 17
experiment # experiment #

Figure 7.9: Numerically calculated and experimental measured data for the (a) 23
solar experimental runs of Campaign 1; and (b) 20 solar experimental runs of Cam-
paign 2, ordered by increasing Q̇solar . Shown are the average cavity and wall tem-
peratures, and the carbon and steam conversions at the outlet. Full circles indicate
numerically calculated values; open circles indicate the experimentally measured
data; the error bars indicate the propagated inaccuracy of the input parameters.

the Al2 O3 liner (eq. (6.9)). Tcavity for the numerical simulation
.P is the weighted mean
P
over all finite volumes of the cavity, Tcavity = Ti Vi ρi Vi ρi , while the lower and
upper error bars describe the 25th and 75th percentile of Ti , respectively. Steam and
7.4. Results and Validation 97

coke conversions are determined from the oxygen and carbon mass balance (Section
3.4.2). The error bars for the calculated values of XC , XH2 O and Tshell are due to
the propagated error of the input parameters, namely ±5% for the coke feeding rate
ṁcoke and ±9%
13% for the solar power input Q̇solar (Section 3.3). The relative RMS
errors between measured and calculated values are indicated (assuming unbiased
input parameters). RMS errors were 5.6%, 20% and 18% for the shell temperature
and for steam and carbon conversion in Campaign 1, and 5.4%, 25% and 39% in
Campaign 2. However, overlapping of the accuracy intervals (error bars) of cal-
culated and measured data points is observed for the majority of the experiments
(error bars for ±5% accuracy for the measured XC and XH2 O not shown in figure).
Finally, calculated temperature ranges in the cavity volume, Tcavity , are found to
be significantly higher than those experimentally estimated (based on the measured
wall temperatures), especially for Campaign 1, in which peak temperatures are ex-
pected to occur at the center of the cavity due to strong radiation absorption by the
small particles.
Table 7.1 lists the selected baseline model parameters of two typical solar ex-

Table 7.1: Baseline model parameters for two representative solar experimental runs
for Campaigns 1 and 2.

Experimental campaign # 1 2
Molar inlet flow rates ṅC , mol/min 0.254 0.102
ṅH2 2 , mol/min 0.142 0.057
ṅH2 O , mol/min 0.446 0.264
ṅAr , mol/min 0.266 0.165
Overall H2 O/C ratio - 1.76 2.6
Inlet velocity uin , m/s 0.065 0.088
Solar power input Q̇solar , kW 5.7 4.7
Static pressure p0 , Pa 105
Inlet equivalent particle diameter D30,in , µm 2.21 17.58
D32,in , µm 3.13 47.14
Reactor cross sectional area S, m2 8 · 10−3
Overall heat transfer coefficient U , W/(m2 K) 12–180
Inlet temperature Tin , K 423.15 1003.5
Surroundings temperature T∞ , K 293.15
2
Bound in feedstock
98 7. Heat and Mass Transfer in the Reactor Cavity

(a)
r = 0.0 m
250 2000
φrad , MW/m3

200
1500

T, K
150 r = 0.025 m

100 1000
φrad
50 Ts
Tg 500
0 {4 {3 {2 {1
10 10 10 10
x, m

(b) 5 2000
φrad
Ts
4 Tg
1500
φrad , MW/m3

3
T, K

r = 0.0 m
2 1000
r = 0.025 m
1
500
0 {4 {3 {2 {1
10 10 10 10
x, m

Figure 7.10: Numerically calculated temperatures profiles for the gaseous and solid
phases and the radiative source term along the reactor at two radial positions: center
(r = 0.0 m) and close to the wall (r = 0.025 m). The baseline parameters listed in
7.1 have been employed for (a) Campaign 1; and (b) Campaign 2.

perimental runs for each experimental campaign presented in Sections 4.1 and 4.2.
Figures 7.10 (a) and (b) show the temperature profiles for the gaseous and solid
phases and the radiative source term along the reactor calculated for the baseline
parameters listed in Table 7.1 at two radial positions: center (r = 0.0 m) and close
7.4. Results and Validation 99

to the wall (r = 0.025 m). Note that the x-axis scale is logarithmic for a bet-
ter appreciation of the rapid heating rate — of the order of 105 K/s — across the
first centimeters after the inlet plane, where particles reach temperatures of 2144
and 1861 K for Campaigns 1 and 2, respectively. Thus, very efficient radiative
heat transfer to the reactants is attained by direct concentrated irradiation of the
gas-particle flow. Temperatures peak after 1 and 5 cm for Campaigns 1 and 2,
respectively, and decrease toward the rear of the reactor as a result of the endother-
mic reaction and conductive heat losses. As expected from the time scales τconv and
τrad , the temperature difference between the solid and gas phases is almost zero for
Campaign 1 with smaller particles, and less than 50 K for Campaign 2 with bigger
particles. The mean temperature of the metallic shell Tshell is in the range 1108–
1301 K and significantly below that of the coke particles because of the Al2 O3 liner
and because the particle suspension serves as a radiation shield, as corroborated
experimentally in a similar gas-particle reactor configuration [110]. The radiation
source term undergoes a fast decrease as the reactants flow along the reactor because
of two phenomena: (1) the absorption coefficient of the particle suspension rapidly
decreases with increasing temperatures due to volumetric expansion and particle
shrinkage as the reaction progresses and (2) the mentioned increase of the particle
temperatures leads to strong emission thus lowering the net radiative source term.
A strong radial temperature gradient is induced by the angular and radial distribu-
tion of the incoming solar radiation. But this gradient disappears after approx. 10
cm behind the aperture as a result of absorption by the particles and emission from
particles and walls.
The progress of the chemical reaction is shown in Figs 7.11 (a) and (b), where the
variation of the chemical composition (chemical species’ molar fractions) is plotted
along the reactor at two radial positions: center (r=0.0 m) and close to the wall
(r=0.025 m). The baseline parameters of Campaigns 1 and 2, listed in Table 7.1,
were employed. Note that the x-scale is shown linear — and not logarithmic as
in Figs 7.10 (a) and (b) — because the chemical reaction rates are significantly
slower than the radiative heat transfer rates. Main product gas components are
H2 and CO, with less than 5% CO2 , as predicted by thermodynamic equilibrium
[101]. Simultaneous fast pyrolysis and steam-gasification occurs immediately after
the entrance of the reactor. The conversion of steam and petcoke is significantly
lower close to the walls (r=0.025 m) as a result of the relatively low temperatures
existing there. In addition, the reaction rates decrease toward the exit of the reactor,
100 7. Heat and Mass Transfer in the Reactor Cavity

(a) 0.5
r = 0.0 m
H2 O r = 0.025 m
0.4

0.3 H2
CO
yi

0.2

0.1 CO2
C

0
0 0.05 0.1 0.15 0.2 0.25
x, m

(b) 0.5
r = 0.0 m
r = 0.025 m
0.4
H2 O
0.3
H2
yi

0.2 C
CO
0.1

CO2
0
0 0.05 0.1 0.15 0.2 0.25
x, m

Figure 7.11: Variation of the chemical composition (chemical species’ molar frac-
tions) along the reactor at two radial positions: center (r = 0.0 m) and close to the
wall (r = 0.025 m). The baseline parameters listed in 7.1 have been employed for
(a) Campaign 1; and (b) Campaign 2.

primarily because of the lower temperatures attained for a decreasing radiation


source, as indicated in Fig. 7.11 (b). As expected, the gasification proceeds at a
higher rate in Campaign 1, due to the higher temperatures and particle effectiveness
η of the feedstock.
7.4. Results and Validation 101

(a) ∂XC , 1/s


∂t
XC
0.05 0.1 0.025 0.01

0.75
0 0.5
0.99
0.25
0.96 0.98
r, m 0.5 0.75 0.9
-0.05
T, K
0.05
2050
1850
1900
0 1950 1800
2000

-0.05
0 0.05 0.1 0.15 0.2 0.25
x, m

(b) ∂XC , 1/s XC


∂t
0.05 0.25 0.15
0.5 0.05
0.75

0 0.95
0.9
0.2 0.6 0.8
r, m

-0.05
T, K
0.05
1250
1000 1700 1650
1725
1600
0

1500
-0.05
0 0.05 0.1 0.15 0.2 0.25
x, m

Figure 7.12: 2D contour map of the carbon conversion XC and reaction rate dXC /dt
(upper plot) and corresponding temperatures (lower plot). The baseline parameters
listed in 7.1 have been employed for (a) Campaign 1; and (b) Campaign 2.

This is also corroborated in the 2D contour maps of Figs 7.12 (a) and (b), showing
the reaction rate dXC /dt and the carbon conversion XC in the upper plot — scaled
with its maximal value — and the corresponding temperature profiles in the lower
plot. In Campaign 1, the peaks for the temperature (> 2050 K) and for the reaction
102 7. Heat and Mass Transfer in the Reactor Cavity

rate 1.4 · 105 1/s are pronounced and located within the first centimeters past the
inlet plane, followed by a decrease to 1800 K and 6 · 10−6 1/s toward the exit of
the reactor. In Campaign 2, a more uniform temperature and reaction rate field is
observed, with maximum values attained 5 cm after the inlet plane. The difference
between the two campaigns is attributed to the differences in the particle sizes of
their feedstock. The smaller particles used in Campaign 1 have a higher absorption
coefficient, as observed from Fig. 5.11, augmenting radiative heat transfer at the
entrance region but preventing radiation to penetrate deeper into the rear of the
cavity. The smaller H2 O/C molar ratio and, consequently, the higher solid volume
fraction, further contribute to this effect. For example, at the location of maximum
temperature the solid volume fractions are 1.83 · 10−5 and 9.75 · 10−6 for Campaigns
1 and 2, respectively, and the corresponding absorption coefficients are 13.3 and 0.28
1/m, respectively. At the exit of the reactor, after condensing the unreacted and
excess H2 O, the molar composition of the syngas was 0.66 H2 / 0.25 CO/ 0.09 CO2
for Campaign 1 and 0.68 H2 / 0.21 CO/ 0.11 CO2 for Campaign 2, which compares
well with the gas composition measured experimentally.

Finally, a sensitivity analysis was performed to elucidate the effect of a given


input parameter Z by computing the system derivative Si = dY /dZ, where Y is
the output of interest [86]. The input parameters are the solar power input Q̇solar ,
the coke feeding rate ṁcoke , the inlet temperature Tin , the emissivity of the walls
wall , and the heat transfer coefficient at the window hw . The outputs of interest
are the mean cavity temperature Tcavity and the coke conversion XC at the reactor
outlet. Results are listed in Table 7.2, where the local relative sensitivity Si,relative
is found by dividing the numerator and denominator of the system derivative Si by
the respective absolute values of Y and Z:
∆Z ∆Z
 
Y Z+ 2
−Y Z − 2
/Y
Si,relative = (7.50)
∆Z/Z

Q̇solar is the input parameter with the strongest impact on the output, especially on
XC for both campaigns. For Campaign 1, increasing ṁcoke has a negative impact
on XC and Treactor . Interestingly, for Campaign 2, its impact on Tcavity is positive
because the feedstock is introduced at a much higher temperature after preheating.
Tin has a positive impact for both campaigns, but stronger for Campaign 2. An
increase in the inlet temperature compensates for the weaker radiation absorption of
7.4. Results and Validation 103

Table 7.2: Relative sensitivity of input/output parameters for both experimental


campaigns.

Campaign Z → Q̇solar ṁcoke Tin wall hw


1 Si,relative (Tcavity ) 0.42 −0.090 0.026 −0.008 −0.004
1 Si,relative (XC ) 1.46 −0.41 0.073 −0.009 −0.003
2 Si,relative (Tcavity ) 0.31 0.014 0.11 −0.015 −0.001
2 Si,relative (XC ) 1.81 −0.16 0.40 −0.11 −0.004

the feedstock used in Campaign 2 because it reduces the energy required to heat the
reactants to the reaction temperature and, consequently, increases the energy left to
drive the chemical reaction. The emissivity of the walls, wall , was found to generally
have a marginal impact on the reactor performance. As expected, the sensitivity to
wall is higher for Campaign 2, because — due to less efficient radiation absorption
by the particles — more radiation is incident directly on the walls. Finally, the
convective heat transfer coefficient at the window, hw , was found to have negligible
impact on the reactor performance.
104 7. Heat and Mass Transfer in the Reactor Cavity
Chapter 8

Optimization and Scale-up1

In the first part of this thesis a solar chemical reactor was designed and a 5 kW
prototype was fabricated and experimentally demonstrated for steam-gasification of
petroleum coke (petcoke) in a high-flux solar furnace. The present chapter focuses
on the optimization and scale-up aspects of such a high-temperature chemical reac-
tor. The tool employed is the two-phase reactor model based on the fundamental
analysis of heat and mass transfer in a directly-irradiated suspension of reacting par-
ticles presented in Chapter 7. Simulations are carried out to determine the optimal
process parameters of the 5 kW prototype reactor and that of the up-scaled 300 kW
reactor. Special emphasis is placed on the feedstock characteristics and the reactor
geometrical configuration, and their effect on the extent of carbon conversion and
solar energy conversion efficiency.

8.1 Simulation Setup


Table 8.1 lists the dimensions and baseline operational conditions of the two solar
reactors analyzed: Synpet5, for a solar power input of 5 kW, and Synpet300 for a
solar power input of 300 kW. Synpet5 was tested at PSI’s solar furnace (Section
3.3); Synpet300 will be tested at CIEMAT’s solar tower at the Plataforma Solar
de Almeria [77]. A scheme of the Synpet300 reactor configuration is depicted in
Fig. 8.1. In contrast to the reactor concept presented in the Chapter 3 the oil cooled
1
Material from this chapter has been submitted for publication as: A. Z’Graggen, and A. Stein-
feld. Hydrogen Production by Steam-Gasification of Carbonaceous Materials using Concentrated
Solar Energy — V. Reactor modeling, optimization, and scale-up. International Journal of Hydro-
gen Energy, 2008.
106 8. Optimization and Scale-up

Table 8.1: Baseline operational parameters for the two reactors analyzed.

Reactor type Synpet5 Synpet300


Cavity volume m3 0.0016 1.4
Aperture diameter Daperture , m 0.05 0.5
Feeding rates ṁcoke , g/min 2.5 500
ṁH2 O , g/min 6.6 1320
V̇Ar , ln/min 4.1 0
Overall molar H2 O/C ratio mol/mol 2 2
Solar power input Q̇solar , kW 5 300
Static pressure p, Pa 105
Overall heat transfer coefficient U , W/(m2 K) 12–180 10
Inlet temperature Tin , K 573.15 773.15
Ambient temperature T∞ , K 293.15

reactor shell (Inconel)


aperture reactor liner
(Al2O3 / SiO2)

tangential injection
nozzles

window purging nozzles product outlet

Figure 8.1: Scheme of the up-scaled chemical reactor configuration for the solar
steam-gasification of coke, featuring a continuous gas-particle vortex flow confined
to a cavity receiver and directly exposed to concentrated solar radiation.

frustum was substituted for a ceramic cone. Furthermore, the thickness of both, the
insulation and the reactor liner were significantly increased to reduce heat losses by
conduction and in order to lower thermal loads on the reactor shell.
The solar spectrum of the incoming radiation is approximated by Planck’s black-
8.2. Results 107

body emission at 5780 K. Figure 5.2 shows the solar power through the aperture as
a function of the aperture’s radius for (a) PSI’s solar furnace, and (b) CIEMAT’s
solar tower. Shown in Figures 5.3 is the angular solar flux distribution at the aper-
ture, used as input to the solar reactor. Spectral values for total transmittance,
reflectance, and absorptance of the 3 mm-thick quartz window are presented in
Fig. 5.7. The overall heat transfer coefficient U , calculated by a FE-solver with
thermal conductivities provided by the manufacturer (Section 5.1.2) ranges from 12
W/m2 K at the center of the cavity to 180 W/m2 K close to the aperture due to
heat bridges for Synpet5, and averages 10 W/m2 K for Synpet300. The inlet tem-
perature Tin is set to 573 K and 773 K for Synpet5 and Synpet300, respectively,
corresponding to an augmented pre-heating of the feedstock for the up-scaled re-
actor. Petrozuata Delayed petroleum coke is employed as a feedstock (Table 2.2).
Most of the optimization calculations were performed with monodisperse particle
clouds with diameters in the range 0.5–1000 µm. In addition, for the purpose of
comparison with the experimental results, three polydisperse particle clouds were
considered: type 1, obtained by subsequent ball and jet milling; type 3, obtained by
80 m-screen sieving; and type 5, the raw feedstock as received from the refinery (see
Table 2.1). Figure 2.2 shows the volume density distribution for the three types of
particles, as determined by laser scattering (Horiba LA950). Mean initial diameters
are 2.21, 17.58 and 40.0 µm for types 1, 3, and 5, respectively. Finally, for the
purpose of computational speedup, convection between the two phases is assumed
instantaneous (Tg =Ts ) and eqs (7.9) and (7.10) simplify to:

∇ · (ρh · u) = φrad (8.1)

In fact, as observed in Fig. 7.10, the temperature difference between the gas and
solid phase vanishes after the initial heating-up, thus having a negligible impact on
the carbon conversion XC at the outlet of the reactor.

8.2 Results
The baseline simulation parameters used for all calculations are listed in Table 8.1.
The effect of varying four parameters was examined: 1) particle size, 2) feeding rate,
3) input power, and 4) geometry of the cavity’s longitudinal section at a constant
cavity volume. The reactor performance is characterized by the extent of feedstock
108 8. Optimization and Scale-up

conversion determined from a carbon mass balance (eq. (3.5)):

ṅCO + ṅCO2 + ṅCH4


XC = (8.2)
ṅ0C

and by the solar-to-chemical energy conversion efficiency ηchem , defined as the portion
of solar energy input stored as chemical energy (eq. (3.7)):

XC · ṅC · ∆HR |298


ηchem = (8.3)
Q̇solar

Particle size Figure 8.2 shows (a) the temperature profiles (average values in the
radial direction), and (b) reaction extents along the reactor axis. The calculation was
performed for the baseline parameters of Synpet5 and with monodisperse particle
clouds of initial diameters 0.57, 0.96, 4.6, 62 and 176 µm, corresponding to curves 1,
2, 3, 4 and 5, respectively. Small particles induce a high volumetric absorption of the
incoming solar radiation, resulting in extremely fast heating rates of 2.9 · 105 K/s,
but at the expense of preventing further penetration of radiation into the cavity.

(a) (b)
2000 1
3
3
1800 0.8 2

1600 4 1
0.6
T, K

XC

2
1400 4
5 0.4
1
1200
0.2 5

1000
0
0 0.1 0.2 0 0.1 0.2
x, m x, m

Figure 8.2: (a) Temperature profiles (average values in the radial direction), and
(b) reaction extents along the reactor axis, for the baseline parameters of Synpet5
(Table 8.1) and with monodisperse particles of initial diameters 0.57, 0.96, 4.6, 62,
and 176 µm, corresponding to curves 1, 2, 3, 4, and 5 respectively.
8.2. Results 109

(a) (b)
ṁcoke /ṁcoke,baseline ṁcoke /ṁcoke,baseline
1 1
0.5
0.5
0.8 1.0 0.8 1.0
0.6 2.0 0.6
XC

XC
2.0
0.4 0.4
4.0
0.2 0.2 4.0
0 type 1 type 3 type 5 0 type 1 type 3 type 5
0.5 1 2 5 10 20 50 200 2 4 10 40 100 400
D, µm D, µm

Figure 8.3: Reaction extent at the exit of the reactor as a function of the feedstock’s
initial particle diameter for (a) Synpet5, and (b) Synpet300. Curves are plotted for
coke and water feeding rates corresponding to 0.5, 1, 2, and 4 times the baseline
values of Table 8.1. Indicated are the equivalent diameters for the polydisperse
feedstock types 1, 3, and 5.

This effect of shadowing is clearly visible for curves 1 and 2. For example, 0.57
µm particles peak at 1920 K in the first centimeter followed by rapid decrease to
1150 K toward the rear of the cavity, leading to a reaction rate slow down. In
contrast, bigger particles absorb the incident radiation only partially, resulting in
slower heating and lower peak temperatures, as observed for curves 4 and 5. For
example, 176 µm particles are heated at 2.1 · 103 K/s and peak at 1670 K after
12 cm from the entrance of the reactor. Curve 3 corresponds to a more uniform
distribution of the radiative absorption over the cavity volume, leading to higher
and more homogenously distributed temperatures, an thus to peak values for XC at
the reactor’s outlet.
Figure 8.3 shows the reaction extent at the exit of the reactor as a function of
the feedstock’s initial particle diameter for (a) Synpet5, and (b) Synpet300. Curves
are plotted for coke and water feeding rates corresponding to 0.5, 1, 2, and 4 times
the baseline values of Table 8.1. For maximum XC , optimal particle diameters
in the range 2–7 µm were obtained for Synpet5, and in the range 11–35 µm for
Synpet300 because of the larger cavity dimensions. Also indicated are the equivalent
110 8. Optimization and Scale-up

diameters D32 for the polydisperse feedstock types 1, 3, and 5. Synpet5 exhibits good
performance with feedstock types 1 and 3, whereas types 3 and 5 are better suited
for Synpet300. Carbon conversions up to 65% and 99% are predicted for Synpet5
and Synpet300, respectively, using feedstock type 5.

Feeding rate Figure 8.4 shows the reaction extent and solar-to-chemical energy
conversion efficiency as a function of the coke mass flow rate (normalized to the
baseline rate) for the three feedstock types 1, 3, and 5, and for (a) Synpet5; and
(b) Synpet300. Also indicated (solid curve) are the results for the optimal initial
particle diameters, as determined in Figure 8.3. In both (a) and (b) cases, complete
reaction extent is achieved for very low feeding rates (ṁcoke /ṁcoke,baseline < 0.39
and < 0.57 for Synpet5 and Synpet300, respectively) at the expense of low energy
conversion efficiency. A moderate increase of the feeding rate to about 1.1 times
ṁcoke,baseline results in an increase in ηchem coupled to a slight decrease in XC . In this
case, the available heat is used efficiently to run the chemical reaction. A further
increase of ṁcoke leads to a drop of the temperatures inside the reactor because of

(a) (b)
optimal D optimal D
1 type 1 1 type 1
type 3 type 3
XC type 5 15 XC type 5
0.8 0.8
25
ηchem, %

ηchem, %

0.6 10 0.6 20
XC

XC

0.4 ηchem 0.4 15


5 10
0.2 0.2
5
ηchem
0 0 0 0
0.5 1 2 4 8 0.5 1 2 4
ṁcoke /ṁcoke,baseline ṁcoke /ṁcoke,baseline

Figure 8.4: Reaction extent (left axis) and solar-to-chemical energy conversion ef-
ficiency (right axis) as a function of the coke mass flow rate (normalized to the
baseline rate) for (a) Synpet5; and (b) Synpet300. The feedstock types are 1, 3, and
5. The solid line is for optimal initial particle diameters in the range 2-7 and 11-35
µm for Synpet5 and Synpet300, respectively.
8.2. Results 111

the heat consumed to heat the reactants, with a consequent decrease of both XC and
ηchem . Optimal feeding rates are in the 0.06–0.4 kg/h range and 17–44 kg/h range
for Synpet5 and Synpet300, respectively. As a consequence of the advantageous
volume-to-surface ratio for the scaled-up reactor and of its lower conduction losses
through a thicker insulation, ηchem can be more than doubled. Values of 10% and
24% are predicted for Synpet5 and Synpet300, respectively.

Input solar power Figure 8.5 shows the (a) solar-to-chemical energy conversion
efficiency, and (b) syngas yield as a function of input solar power for Synpet300 using
feedstock type 5. The syngas composition is characterized by H2 /CO molar ratios
in the range 2.6–9.7 and CO2 /CO molar ratios in the range 0.6–3.2. The dotted,
dash-dotted, and dashed lines represent the performance for selected reaction extents
of 0.8, 0.95, and 0.99, respectively. The solid curve in both figures represents the
locus of maximum ηchem and the corresponding reaction extents (Fig. 8.5a) and
petcoke feeding rates (Figure 8.5b). Evidently, higher XC is obtained at the cost of

(a) (b)
60 72
0.77 ṁcoke , kg/h
XC 0.75
25
0.71 50
0.64 58
20 CO
40
ηchem, %

ṁ, kg/h

15 0.52 30 45
10 20 H2
optimal
XC = 0.8 31
5 XC = 0.95 10
17
XC = 0.99
0 0
200 300 400 500 600 200 300 400 500 600
Q̇solar , kW Q̇solar , kW

Figure 8.5: Solar-to-chemical energy conversion efficiency (a), and syngas yield (b),
as a function of input solar power for Synpet300 using feedstock type 5. The dotted,
dash-dotted, and dashed lines represent the performance for selected reaction extents
of 0.8, 0.95 and 0.99, respectively. The solid curve in both figures represents the
locus of maximum ηchem . Indicated are selected corresponding values of XC in (a)
and ṁcoke in (b).
112 8. Optimization and Scale-up

lower ηchem . For example, for Q̇solar =464 kW (spring equinox), ṁcoke =35 kg/h, and
XC =95% yields ṁH2 =9.1 kg/h and ṁCO =39.8 kg/h at ηchem =20% (H2 /CO=3.2 and
CO2 /CO=0.8). An analogous calculation performed for feedstock type 3 yielded, as
expected, superior performance: ηchem =24%, ṁH2 =10.8 kg/h and ṁCO =46.4 kg/h
for ṁcoke =41 kg/h (H2 /CO=3.3 and CO2 /CO=0.9). However, feedstock type 3
requires an additional processing step.

Geometry The seven cavity cross-sections shown in Fig. 8.6 were examined for
(a) Synpet5 and (b) Synpet300, where the radius and the length of the reactor are
varied while the cavity volume is kept constant; The geometries are labeled from
‘–3’ to ‘3’, corresponding to radii from 9.4 to 2.5 cm and overall reactor lengths
from 9.5 to 90 cm for Synpet5, and to radii from 121 to 25 cm and overall reactor
lengths from 71 to 686 cm for Synpet300. ‘0’ corresponds to the baseline design
shape of Synpet5 and Synpet300. Figure 8.7 shows the contour map of reaction
extents as a function of geometry and particle diameter for the baseline parameters
listed in Table 8.1 and for (a) Synpet5 and (b) Synpet300. Interestingly, Fig. 8.7a

(a)
–3 –2
0.1 –1 0 1 2 3
r, m

0
–0.1
0 0.2 0.4 0.6 0.8
x, m

(b)
–3
1 –2
–1 0
1 2 3
r, m

–1

0 1 2 3 4 5 6 7
x, m

Figure 8.6: Reactor cross-sections of constant volume where geometry ‘0’ corre-
sponds to the baseline dimensions (Table 8.1).
8.2. Results 113

(a)
50
0.6
0.75
20
10 0.9 0.85
D, µm 5 0.9
XC 0.94
2 0.85
1
0.4 0.6 0.75
0.5

–3 –2 –1 0 1 2 3
geometry #

(b)
100
0.5 0.75
50 0.85
0.9
20
D, µm

0.25 0.94
10
5
0.1
2 XC
1
–3 –2 –1 0 1 2 3
geometry #

Figure 8.7: Contour maps of reaction extent as a function of geometry (see Fig. 8.6)
and particle diameter for the baseline parameters listed in Table 8.1 and for: (a)
Synpet5, and (b) Synpet300.

shows two local maxima. The first, located at D = 5 µm for geometry #–1, is due
to a minimum in the conductive heat losses — 56% compared to a maximum of
62% of Q̇solar —, while the radiative losses stay constant at 19% of Q̇solar . Since
geometry #–1 approaches a sphere, it has the smallest surface-to-volume ratio. A
second maximum is found at D = 2 µm for geometry #3, as the cavity’s radius
114 8. Optimization and Scale-up

approaches the aperture’s radius. In this case, all injected particles are exposed to
the peak flux and reach up to 2090 K compared to 1922 K for the first maximum.
As a consequence, the chemical reaction rate is faster and takes place in the 1st
half of the reactor. Conduction losses become less important (37% of Q̇solar ) due to
the somewhat lower wall temperatures in the 2n d half of the reactor, as products
exit at 582 K compared to 1743 K for the first maximum. In contrast, re-radiation
losses increase dramatically (48% of Q̇solar ) due to the high temperatures reached
in the front of the cavity. Figure 8.7b shows one maximum located at D = 8 µm
between geometry #2 and #3. In contrast to the calculations performed for Synpet5
(Fig. 8.7a), Synpet300 does not show a second maximum as it approaches a spherical
geometry because, due to the increased insulation thickness, conduction losses play
a lesser role.
Chapter 9

Conclusions

This thesis work was performed in the framework of a joint project between the
Swiss Federal Institute of Technology in Zurich (ETH) and the Paul Scherrer Insti-
tut in Villigen (PSI), both in Switzerland, the national research center ‘Centro de
Investigaciones Energéticas, Medioambientales y Tecnológicas’ in Spain (CIEMAT)
and the research and development center of Petróleos de Venezuela, S.A. (PDVSA
/ INTEVEP). The main goal of the still-ongoing project is the development and
demonstration of a process and of the related technology required for the produc-
tion of high quality syngas from carbonaceous materials using a solar thermochemical
process. ETH’s and PSI’s working packages included six tasks: (1) chemical thermo-
dynamics analysis, (2) second-law (exergy) analysis, (3) chemical kinetics analysis,
(4) engineering design and operation of a 5 kW prototype reactor (Synpet5), (5)
engineering design of the scaled-up 300 kW reactor (Synpet300) and (6) heat and
mass transfer modeling. Tasks (1)–(3) were performed by D. Trommer in a first PhD
Thesis [100], whereas tasks (4)–(6) were treated in the present thesis; main foci were
the Synpet5 experimental results and the modeling of both, Synpet5 and Synpet300.
The engineering design, fabrication, and testing of both reactors was the result of a
joint effort at the Professorship of Renewable Energy Carriers in cooperation with
the project partners.

9.1 Experimental Work


A solar chemical reactor for performing solar steam-gasification of carbonaceous
materials to syngas was successfully designed and tested. Solar experiments in a
116 9. Conclusions

high-flux solar furnace were performed for three distinct feedstock types. Coke
particles where either fed as dry powder (mean particle diameter 2.2 µm) or mixed
with water to form a slurry (mean particle diameter in the range 8.5–200 µm).
Vacuum residue was liquefied prior to injection. The reactor was operated smoothly
for approximately 100 experiments, each one lasting around one hour, yielding 64
valid experimental runs presented in Part I of this thesis. Best reactor performances
were measured for dry coke feeding, with carbon conversions of up to 87%, solar-
to-chemical energy conversion efficiencies up to 9%, typical reactor temperatures
above 1600 K and mean residence times around 1 second. The somewhat lower
values obtained for slurry and vacuum residue were caused by non-optimal feeding
systems, because the original reactor was designed for dry coke powder, not by
intrinsic drawbacks of these feedstock types. In fact, the design of the scaled-up
reactor foresees the use of coke-water slurries, which will facilitate feedstock handling
and control at large scales. The quality of the syngas produced was comparable to
the one typically obtained when heat is supplied by internal combustion. In general,
results indicate the technical feasibility of pyrolysing and steam-gasifying various
carbonaceous materials with concentrated solar energy.

9.2 Heat and Mass Transfer Modeling


A simple lumped-parameters model was formulated for solving the steady-state en-
ergy conservation equation that links the radiative power input with the power
consumed by the endothermic chemical reaction. The simulation was used, together
with the thermodynamic and kinetic analyses, to support the engineering design of
the Synpet5 reactor as well as to estimate the up-scaling potential. The retrospective
comparison the the experimental runs showed good agreement for the reactor’s tem-
peratures. Nevertheless the lumped-parameters model was not able to accurately
reproduce chemical conversion of carbon and steam at the reactor’s exit, especially
for Campaign 2.
A more sophisticated model was therefore developed which accounts for heat
and mass transfer in a polydisperse particle suspension subjected to concentrated
thermal radiation and undergoing an endothermic chemical transformation. The
exact reactor’s geometry was implemented while the incoming solar radiation was
characterized with preliminary MC calculations and the conduction in the reactor
walls was predicted by a FE simulation. Numerical results for average temperatures,
9.3. Outlook 117

coke and water conversion, and gas composition at the reactor outlet agreed well
with those obtained in the two experimental campaigns run with coke. The differ-
ent particle size distributions of the feedstock used in the two campaigns strongly
affected the absorption coefficient and chemical kinetics, and consequently mass and
heat transfer, justifying the detailed analysis of the polydisperse medium.
Finally the latter model was use to examine the lab-scale 5 kW prototype (Syn-
pet5), and its scaled-up version of 300 kW (Synpet300). The feedstock’s particle
size, the feeding flow rates, the solar power input, and the geometry of the reactors
were varied to identify the optimal operational conditions for maximum solar-to-
chemical energy conversion efficiency. Optimal initial feedstock particle sizes were
in the range 2–7 and 11–35 µm for the 5 and 300 kW reactors, respectively. The ad-
vantageous volume-to-surface ratio of the 300 kW scaled-up reactor and its enhanced
insulation lead to an optimal coke feeding rate of 44 kg/h for an solar-to-chemical
energy conversion efficiency of 24%, resulting in a syngas yield of ṁH2 =8.8 kg/h
and ṁCO =15.6 kg/h. Moreover, the increased cavity dimensions permit the use of a
coarser feedstock to be reacted efficiently. Last, but not least, the tube-shaped cav-
ity was found to outperform the commonly employed barrel-shaped cavity, mainly
because most of the particles are directly exposed to the incoming high-flux solar
radiation.

9.3 Outlook
The scaled-up reactor designed by the ETH-team has been fabricated by CIEMAT
in 2007. The complete pilot plant is currently being assembled at the Plataforma
Solar de Almerı́a in Spain, where the Synpet300 reactor will be installed on top of
the SSPS central receiver system. The new reactor is scheduled for operation in
2008 and will mark a further milestone in the development of a new gasification
technology for the solar upgrading of low-valued carbonaceous materials. The ulti-
mate objective is the commercial implementation of a solar gasification plant with
multiple receiver/mirror fields for the operation on an industrial MW scale.
The detailed model used to simulate radiative heat transfer in polydisperse re-
acting particulate media presented in this thesis was coupled to a rather simple fluid
flow solver, as the specific operational parameters allowed several simplifications. In
a more general context it is desirable to couple the radiation solver (here developed)
to a common 3D CFD full-featured solver, as these solvers typically lack of accu-
118 9. Conclusions

rate models for the treatment of participating media. In addition, model validation
would be enhanced by improving the accuracy of the experimental input parameters
such as Q̇solar and ṁfeedstock , and by measuring also local flow properties such as gas
temperature and chemical composition in the volume. A measurement device for the
determination of gas temperatures in highly radiating environments was developed
for this purpose (Appendix A).
Part III

Appendices
Appendix A

Gas Temperature Measurement in


Radiating Environments1

An experimental methodology is developed for gas temperature measurements in


highly radiating environments. It consists of a suction thermocouple apparatus and
associated heat transfer model for determining the gas temperature from shielded
thermocouple readings by radiation, convection, and conduction dimensionless cor-
relations. The apparatus and methodology are calibrated and applied to measure
gas flow temperatures in a tubular furnace with wall temperatures up to 1223 K.
Results are compared with predictions by CFD simulations.

A.1 Additional Nomenclature

Latin Characters
ci parameter for empirical correlations
D diameter, m
L length, m
Greek Characters
δ relative difference
Subscripts
bare unshielded thermocouple
1
This chapter has been published as: A. Z’Graggen, H. Friess, and A. Steinfeld. Gas temper-
ature measurement in thermal radiating environments using a suction thermocouple apparatus.
Measurement Science and Technology, 18:3329–3334, 2007.
122 A. Gas Temperature Measurement in Radiating Environments

furnace furnace
gas gas
Mo molybdenum, material of the thermocouple mantle
film gas film
sh shield
suc suction
tc shielded thermocouple
Dimensionless Numbers
Bi Biot number, h · L/k
Gz Graetz number, Re · Pr · D/L
Nu Nusselt number, h · D/kgas
Pr Prandtl number, µ · cp /k
Re Reynolds number, ρ · u · D/µ

A.2 Introduction
Heat and mass transfer in radiating environments is of importance in numerous
high-temperature applications. Examples are combustion flames [108], fluidized
beds [118, 55], coal-fired furnaces [32, 64, 61], and, more recently, solar thermo-
chemical processes [60, 40, 71]. Especially the analysis of radiation heat transfer in
absorbing-emitting-scattering media is often based on model approximations. The
experimental validation of the resulting temperatures in radiative equilibrium plays
therefore a fundamental role in assessing the quality of such models. Optical tem-
perature measurements have been widely applied [17], but these require optical ac-
cess through protected transparent windows, which introduce design complications.
Fluid temperatures are commonly measured by means of bare thermocouples, but
measurement errors are caused by conduction along the thermocouple and by ra-
diative exchange with the surroundings [66]. Above 1000 K, the error induced by
radiation can be of several hundred degrees when the temperature difference between
fluid and surroundings is relative large and/or when the thermocouple is not prop-
erly shielded from incoming radiation [9]. The error can be significantly reduced by
making use of shielded suction thermocouples, where the shield attenuates the detri-
mental effect of radiative heat transfer and the suction flow amplifies the beneficial
effect of convective heat transfer. Suction - also referred to as aspiration - thermocou-
A.3. Experimental Setup 123

ples have been analyzed as a function of number of shields, aspiration velocity, and
surroundings temperature [9]. Correct temperature readings can then be achieved
by increasing the suction speed until no further change is observed [50], by correction
models based on intermittent measurements [39], and by correction models based on
the estimation of the radiative and convective heat transfer [63, 116]. The aforemen-
tioned studies are limited to their specific application because they suffer from one
or more of the following constraints: (1) information about the surrounding temper-
atures is required but difficult to estimate; (2) the coefficients for convective heat
transfer are affected by high uncertainties; and, (3) relatively large suction volumes
are needed. The present study shows how to alleviate these problems and provides
a more general methodology for the estimation of the measurement accuracy which
accounts for convection, radiation, and conduction heat transfer. The apparatus
is calibrated and applied to measure gas flow temperatures in a tubular furnace
with wall temperatures in the range 623-1223 K. A corroboration is accomplished
by comparison with predictions by CFD simulations.

A.3 Experimental Setup


The experimental setup is schematically shown in A.1. It encompasses two config-
urations for measuring the gas temperatures in a furnace and for calibrating the
measurement apparatus. A close-up view of the suction thermocouple apparatus
(lance) is depicted in A.2. It is composed of a 1 mm diameter (Dtc ) thermocou-
ple type K, shielded by an Inconel 600 tube of 4 mm inner diameter (Dsh ), 8 mm
outer diameter, and 30 mm length. Two additional thermocouples type K are built
into the shield tube at the same axial position. The shield is further attached to a
supporting Inconel 600 tube of 8 mm inner diameter, 10 mm outer diameter, and
200 mm length. Gas is aspired through the probe by means of a membrane vacuum
pump, while pulsations are attenuated by a 0.75 liters steel pressure vessel. Down-
stream, the gas flow is cooled to ambient temperature in a heat exchanger (HX)
and passes through an electronic flow controller (Vögtlin Instruments: red-y smart
series). The measurement apparatus is mounted inside a tubular furnace (Carbolite
MTF 10/25/130 with a max. nominal temperature of 1300 K, equipped with a 18
mm inner diameter, 300 mm length Al2 O3 working tube). From the opposite side,
gas is injected into the furnace at a controlled mass flow rate. The flow pattern is
rectified by means of a ceramic honeycomb structure. At the exit of the furnace,
124 A. Gas Temperature Measurement in Radiating Environments

measuring unit data aquisition suction unit

V̇ furnace Tfurnace membrane


Ttc
pump
Tsh1
Tsh2
pulsation
attenuator
lance V̇ suc
flow furnace
N2 rectifier
HX flowcontroller
heating coil
air @ Tambient
lance
calibration unit

Figure A.1: Scheme of the experimental setup.

Lin Ltc A
Tbare thermocouples

Dsh Dtc
Tgas V̇suc

shield A
V̇in
Tsh1
Ttc
Tbare
Q̇rad Tsh2
section A-A

Figure A.2: Scheme of the suction thermocouple apparatus.

the gas is released to the environment via a cooling tower. For the purpose of cali-
bration, the measurement apparatus is removed from the furnace and an electrical
heating coil is applied to the first 100 mm of the lance to simulate a strong radiative
environment.
A.4. Heat Transfer Model 125

A.4 Heat Transfer Model

The numerical heat transfer model is based on the application of the energy conser-
vation equation to the tip of the shielded thermocouple. The domain is indicated by
the dashed box in A.2. Three heat transfer modes are considered: 1) convection from
the shield to the gas and from the thermocouple to the gas, 2) radiation between the
thermocouple and the shield, and 3) conduction at the base of the thermocouple.
The Bi number in the radial direction at the tip of the thermocouple,Bi = h·Dtc /kMo ,
has values around 10−3 , justifying the lumped parameter model for the thermocou-
ple temperature Ttc . For flow regimes with Gzsh > Gzlim , where Gzsh = DLinsh · Resh · Pr
and Gzlim ≈ 20 , the tip of the probe is in the thermal entrance region and con-
vection from the shield can be neglected [44]. The convective heat transfer at the
thermocouple tip can therefore be expressed by:

Q̇conv = Stc · h · (Tgas − Ttc ) (A.1)

where Stc is the thermocouple’s surface exposed to the gas flow and is the av-
erage heat transfer coefficient calculated from h = Nu · kgas (Tfilm ) /Dtc , evalu-
ated at the temperature of the gas film surrounding the thermocouple tip, Tfilm =
0.5·(Tgas + Ttc ). The mean Nusselt number is derived from the empirical correlation:

Nu = c1 · Rectc2 · Pr1/3 (A.2)

where the Reynolds number at the tip of the thermocouple is given by:

u · ρ (Tfilm ) · Dtc 4 · Dtc ṁsuc


Retc = =  · (A.3)
µ (Tfilm ) 2
π · Dsh−D 2
µ (Tfilm )
tc

The view factor from the thermocouple to the gas inlet is less than 2%. Thus,
the influence of radiation from the surroundings is neglected. The radiative heat
transfer is then estimated using the expression for diffuse radiative exchange between
concentric cylinders:
4
Stc · σ · (Tsh − T4)
Q̇rad =   tc (A.4)
1 Dtc 1
tc
+ Dsh
· sh
− 1
126 A. Gas Temperature Measurement in Radiating Environments

where Tsh = 0.5·(Tsh1 + Tsh2 ) . Since the base of the thermocouple is attached to the
shield, the shield temperature is included in the conductive heat transfer calculation:

Tgas − Ttc
Q̇cond = keff · Scond · (A.5)
Ltc

where Ltc is the characteristic length, Scond is the thermocouple cross section, and
keff is the effective thermal conductivity given by the empirical correlation,
 c4
Tsh − Ttc
keff = kMo (Ttc ) · c3 · (A.6)
Tsh − Tgas

The parameter c3 accounts for temperature-independent effects, in particular for


the fact that the whole thermocouple cross-section is not relevant for conduction
since its core is made of insulating SiO2 . The parameter c4 accounts for nonlinear
effects, such as convective heat transfer from the gas to the shield at the clamping
of the thermocouple and nonlinear temperature profiles. The parameters c1 and c2
of eq. (A.2), and c3 and c4 of eq. (A.6) are identified by a set of calibration measure-
ments performed with air (see Section A.5). Heat balance at the thermocouple tip
yields:
Q̇cond + Q̇conv + Q̇rad = 0 (A.7)

Finally, the gas temperature is calculated by the implicit equation:

1
Tgas = Ttc − · Q̇rad + Q̇cond (Tgas ) (A.8)
Stc · h̄

using the Nelder-Mead Simplex method [69].

A.5 Calibration Measurements


The experimental operating conditions and results for a set of calibration measure-
ments performed with air are listed in Table A.2. Geometrical dimensions and
material properties are listed in Table A.3.
Three sets of calibration runs were performed with nominal shield temperatures
of 473, 573, and 673 K. The suction flow was varied between 0.5 and 2.9 ln /min, re-
sulting in Re at the thermocouple tip in the range 35-257 (eq. (A.3)). The reciprocal
Graetz number was always smaller the 0.05, namely in the range 0.0036-0.0214. For
A.5. Calibration Measurements 127

Table A.2: Experimental operating conditions and measurement results for the cal-
ibration of the suction thermocouple apparatus.

V̇suc ṁsuc · 106 Ttc Tsh Tgas Retc 1/Gzsh


ln min−1 kg s−1 K K K – –
0.49 9.56 524 674 293 35 0.0214
0.99 19.13 435 673 293 76 0.0107
1.48 28.71 403 674 293 119 0.0071
1.98 38.29 382 673 293 162 0.0053
2.47 47.86 365 653 293 207 0.0043
2.81 54.45 354 636 293 238 0.0038
0.49 9.56 453 575 293 37 0.0214
0.99 19.13 388 572 293 80 0.0107
1.48 28.71 362 572 293 124 0.0071
1.98 38.29 350 574 293 168 0.0053
2.47 47.86 344 574 293 212 0.0043
2.81 54.45 340 575 293 242 0.0038
0.49 9.58 387 473 293 40 0.0214
0.99 19.15 348 474 293 84 0.0107
1.48 28.73 331 471 293 129 0.0071
1.98 38.31 324 476 293 174 0.0053
2.47 47.86 320 472 293 218 0.0043
2.91 56.31 317 471 293 257 0.0036

example, a suction flow of 1.48 ln /min at an imposed shield temperature of 471 K


resulted in a thermocouple reading of 331 K and values of 179.5 W/(m2 K) and 15.3
W/(m K) for the heat transfer coefficient and the effective thermal conductivity,
respectively. Equations (A.2) and (A.6) were fitted to the experimental data by
nonlinear regression using the least squares method. The results of the parametric
identification for c1 , c2 , c3 , and c4 , and the accuracy of the regression R2 are listed in
Table A.4. Figure A.3 shows the comparison of the measured Nusselt numbers with
the empirical correlation. The bars show the propagated error for a measurement
accuracy of ±2 K for temperatures and ±1.5% for suction mass flow rates. The
ratio Gzsh /Gzlim is found to be larger than 1 for every measurement, implying that
the tip of the thermocouple belongs to the entrance region. The calibrated model
can be applied to gas species and temperatures beyond the calibration conditions
thanks to the dimensionless nature of the heat transfer model.
128 A. Gas Temperature Measurement in Radiating Environments

Table A.3: Geometrical dimensions of the suction thermocouple apparatus and ma-
terial properties for air, N2 , and Mo.

Lin m 0.01
Ltc m 0.01
Dsh m 0.004
Dtc m 0.001
sh 2 0.8
tc 2 0.8
kMo 3 W/(m · K) 138–104
kAir 4 W/(m · K) 0.024–0.084
µAir 4 kg/(m · s) 1.8 · 10−5 –4.8 · 10−5
kN2 4 W/(m · K) 0.025–0.077
µN2 4 kg/(m · s) 1.7 · 10−5 –4.5 · 10−5
2
Oxidized surfaces assumed gray for Inconel [117] and Molybdenum [99].
3
Values presented for 293 K and 1223 K [98].
4
Values presented for 293 K and 1223 K [106].

Table A.4: Parameters for the empirical correlations of the Nusselt number (c1 and
c2 in eq. (A.2)), and of the effective thermal conductivity (c3 and c4 in eq. (A.6),
and regression accuracy R2 .

c1 c2 c3 c4 R2
0.2867 0.6806 0.0779 –1.4973 0.993

A.6 Furnace Measurements


The suction thermocouple apparatus and associated methodology were applied for
determining gas temperatures of a N2 flow through a tubular furnace at different
furnace temperatures. The experimental operating conditions, measurement results,
and calculated gas temperatures are listed in Table A.5. Experiments were carried
out at three values of temperature, Tfurncae = 623, 923 and 1223 K. Nitrogen gas flow
rate to the furnace varied between 5 and 10 ln /min, whereas the suction flow was
in the range 0.5-2 ln /min. Gas temperatures were calculated using eq. (A.8) with
parameters of Table A.4. Figure A.4 shows the experimental data (solid lines) and
calculated gas temperatures (dashed lines) as a function of suction mass flow rate for
1223 K and a N2 flow to the furnace of 10 ln /min. The temperature of the shield Tsh ,
A.7. CFD Simulations 129

15
0.2867 · Re0.6806
tc · Pr1/3
10 Nu
Gzsh /Gzlim

Gzsh /Gzlim
16
5
Nu

3 4

2
2
1
30 50 100 200 300
Retc

Figure A.3: Left axis: experimentally determined and numerically computed Nus-
selt number as a function of Reynolds number. Right axis: corresponding Graetz
numbers, normalized by Gzlim .

as well as the temperature of the bare thermocouple Tbare , is almost independent of


the suction flow since the convective heat transfer plays a minor role. In contrast,
and as expected, the temperature of the central thermocouple Ttc strongly depends
on Vsuc , approaching the gas temperature with increasing suction flow. For example,
for Tfurnace= 1223 K, the average temperatures for Tsh , Tbare , Tgas and are 993, 889,
and 452 K, respectively. The shielded thermocouple reading Ttc was 829, 733, 668
and 634 K for suction flows Vsuc of 0.6, 1.0, 1.5 and 2.0 ln /min, respectively. Also
indicated in Figure A.4 is the gas temperature TgasCFD obtained by CFD simulations
(see Section A.7).

A.7 CFD Simulations


The gas temperature Tgas determined using the suction thermocouple apparatus
was compared with that obtained by CFD simulations TgasCFD . Their difference
can be considered as a reasonable estimate of the accuracy of the measurement
methodology. The computational domain is limited to the gas volume inside the
working tube of the furnace. The simulated gas is N2 with properties taken from
[106] and it is assumed to be non-participating with respect to radiation. Reynolds
130 A. Gas Temperature Measurement in Radiating Environments

Table A.5: Experimental operating conditions, measurement results, and calculated


gas temperatures with N2 at different furnace temperatures.

V̇furnace ṁfurnace V̇suc ṁsuc Tfurnace Tbare Ttc Tsh Tgas TgasCFD Retc 1/Gzsh
ln /min mg/s ln /min mg/s K K K K K K – –
5.00 93.67 0.60 11.24 623 463 451 517 406 412 41 0.0222
5.00 93.67 1.00 18.73 623 460 436 514 406 423 69 0.0133
10.00 187.34 0.50 9.37 923 609 567 708 387 352 32 0.0258
10.00 187.34 1.00 18.73 923 605 495 691 391 354 67 0.0130
10.00 187.34 1.50 28.10 923 602 471 682 395 357 102 0.0087
10.00 187.34 2.00 37.47 923 598 459 676 399 361 137 0.0066
10.00 187.34 0.50 9.37 1223 908 867 1021 435 398 26 0.0278
10.00 187.34 0.60 11.24 1223 906 829 1015 448 398 32 0.0237
10.00 187.34 1.00 18.73 1223 901 733 999 450 397 55 0.0143
10.00 187.34 1.50 28.10 1223 896 668 985 451 398 86 0.0095
10.00 187.34 2.00 37.47 1223 893 634 977 458 401 116 0.0072

numbers obtained are in the range 149-742, justifying the laminar flow assumption.
The following boundary conditions (BC) are set:

• working tube wall : imposed temperature determined experimentally (see Table


A.6);

• shield : imposed temperature equal to Tsh ;

• thermocouple: imposed temperature equal to Ttc ;

• furnace inlet: plug flow with imposed mass flow rate ṁfurnace at ambient tem-
perature 297 K;

• furnace outlet: imposed pressure equal to 1 bar;

• thermocouple outlet: imposed mass flow rate equal to ṁsuc .

Since gravity causes the shape of flow field to differ significantly from rotational
symmetry, only one symmetry plane is employed, which divides the domain in two
halves. The simulations are based on tetrahedral discretization meshes for one of
these half-domains. The discretization error for the computed temperatures was
found to be less than 1 K by stepwise grid refinement. The tabulated values of
TgasCFD have been derived from the raw simulation results by averaging over a circular
A.7. CFD Simulations 131

Tbare Ttc TgasCFD


Tsh Tgas
1100

950
T, K
800

650

500

350
0.5 1 1.5 2
V̇suc , ln /min

Figure A.4: Experimental data (solid lines) and calculated gas temperatures (dashed
lines) as a function of the suction mass flow rate at a nominal furnace temperature
of 1223 and N2 mass flow rate of 10 ln /min.

disk located at the axial position of the lance inlet (195 mm from the working tube
inlet). The diameter of the disk is set equal to 0.75 · Dsh , ensuring that TgasCFD is
not affected by the temperature boundary layer belonging to the radiation shield.
The following relative temperature differences are defined:

Tbare − TgasCFD
δbare = (A.9)
TgasCFD

Ttc − TgasCFD
δtc = (A.10)
TgasCFD

Tgas − TgasCFD
δmethod = (A.11)
TgasCFD

Figure A.5 shows δbare , δtc , and δmethod as a function of the temperature of the
surroundings (Tfurncae ) and of Re at the thermocouple tip. As expected, two impor-
tant effects can be observed. Firstly, high temperatures have a detrimental effect
on the readings of the bare and the shielded thermocouples as a result of radia-
tion heat transfer. Secondly, high Re numbers have a beneficial effect on Ttc as a
result of convective heat transfer. In fact, for the limiting case of Retc → ∞, Ttc
approaches the gas temperature. The relative difference between the temperature
132 A. Gas Temperature Measurement in Radiating Environments

Table A.6: Measured temperatures of the working tube in the furnace as a function
of the axial position x, starting from the inlet. The center of the furnace as well as
the entrance of the suction thermocouple apparatus is at position x = 0.195 m.

x, m
Tfurnace , K 0.0 0.03 0.06 0.09 0.12 0.15 0.18 0.21 0.24 0.27 0.3
TBC , K
1223 340 340 366 425 622 1082 1278 1301 1247 773 662
923 311 325 350 400 515 846 966 981 929 653 503
623 311 321 337 370 432 555 637 663 646 513 411

0
10

120 Retc
40 80
δ

–1
10

δbare
δtc
δmethod
–2
10
650 775 900 1025 1150
Tfurnace , K

Figure A.5: Relative differences between the gas temperature simulated by CFD
(TgasCFD ) and the bare thermocouple reading (Tbare ), the shielded thermocouple read-
ing (Ttc ), and calculated gas temperature (Tgas ), at different Reynolds numbers. See
definitions of δbare , δtc , and δmethod in equations (A.9), (A.10), and (A.11), respec-
tively.

measured with the bare thermocouple and that of the gas is 12, 70, and 126% for
furnace temperatures of 623, 923, and 1223 K, respectively. For increased suction
flows, the difference is somewhat lower. For example, for Tfurnace = 923 K, δtc = 55,
36, and 29% for Re = 40, 80, and 120, respectively. The relative difference between
Tgas and TgasCFD varies from 2% to 10% for Tfurnace from 623 to 1223 K, as a result
of the approximations in the boundary conditions used for the CFD simulations.
A.8. Summary and Conclusions 133

A.8 Summary and Conclusions


An apparatus and associated methodology for gas temperature measurements in
highly radiating environments has been developed. The main advantages of the
presented approach are threefold: 1) knowledge of the surroundings’ temperature
is not required; 2) it is applicable to gas mixtures and temperatures beyond the
calibration conditions; 3) small suction mass flow rates (∼ 0.5 ln /min) are sufficient,
minimizing the impact on the experimental environment. The relative large differ-
ences between the true gas temperature and the readings of the bare and shielded
thermocouples demonstrate the importance of an accurate and validated heat trans-
fer model. Miniaturization of the measurement apparatus would further enhance the
accuracy at a given suction flow due to larger Reynolds numbers. In addition, the
resulting lower thermal inertia would lead to faster response, and it would reduce
the influence of the external flow pattern on the flow inside the apparatus.
134 A. Gas Temperature Measurement in Radiating Environments
Appendix B

Measurement Accuracy of Q̇solar

The PSI solar concentrator does not permit the continuous online measurement of
the solar power entering the reactor trough its aperture, Q̇solar . The position of
the Venetian-blind type shutter ηshutter , defined as the ratio between the radiation
outside and behind the shutter, and solar radiation incident on the heliostat Isolar are
recorded instead. Solar power input is then calculated with the following empirical
correlation:
Q̇solar = Isolar · ηshutter · k (B.1)

where k is found from measurements taken before and after the experiments. There,
the image of the concentrated solar radiation on a diffusely reflecting target is ana-
lyzed by a calibrated CCD camera, resulting in an estimate for Q̇solar . The location
of the maximal incoming power is then identified manually and the reactor setup is
moved until this location and the aperture coincide. There are two main sources of
error in this procedure: (1) the values of k are strongly influenced by soiling of the
heliostat, the parabolic concentrator and the target. Furthermore, k is also affected
by the care of the operator that estimates Q̇solar . Typical average values for k of 8.5,
6.9 and 7.9 with a standard deviation of 12.9%, 7.9% and 7.6% were obtained for

Table B.1: Influence of the inaccurate positioning of the aperture with respect to
the focus of the parabolic concentrator (spot).

placement error, m 0 0.005 0.01 0.02 0.03 0.04 0.05


Q̇solar , kW 5.0 4.97 4.79 4.14 3.24 2.31 1.51
relative error, % 0 0.52 4.28 17.26 35.11 53.78 69.76
136 B. Measurement Accuracy of Q̇solar

Campaign 1, Campaign 2 and Campaign 3, respectively. (2) the reactor’s aperture is


not perfectly aligned to the location of the maximal power (spot). A slow dislocation
of the spot was observed during operation as a consequence of imperfect tracking
of the sun path by the heliostat. The impact of this misalignment on Q̇solar was
analyzed with the same simulation setup used in Section 5.1.1 to characterize the
angular and radial distribution of the incoming solar radiation. Results are shown
in Table B.1. A placement error of up to 5 cm was investigated. During typically
operation the bias does not exceed 1 cm, producing an error of ±04.3 %. This error
combined with the average standard deviation of k then results in an estimated
accuracy of ±913 % for Q̇solar . Note that additional error caused by occasional dust
deposition on the window of the reactor is not taken into account.
The issue of online flux measurements on solar furnaces has been addressed by
several research groups. Interesting approaches, including the ‘moving bar’ principle,
are found in [5] and [104].
Appendix C

Radiation Absorption in the Gas


Phase1

The absorption coefficient of a single spectral line for a gas at temperature T , partial
pressure p and at wavenumber ν is described by [83]:

aij (ν, T, p) = Sij (T ) f (ν, νij , T, p) (C.1)

where Sij is the temperature-dependent line intensity, f is the normalized line shape
function, which takes into account temperature and pressure correction of the line
halfwidth as well as pressure shift of the line position, and νij is the transition
wavenumber of the line considered. The spectral absorption coefficient of a gas is
then calculated as the sum of all contributions aij at wavelength λ for each line:

pNA X
κλ = aij (1/λ, T, p) (C.2)
RT lines

Figure C.1 shows results for high resolution calculations in the range 0.1-40 µm
performed for a gas mixture of 50% CO and 50% H2O and at 1800 K and 1 bar for
the 918 (CO in HITEMP) and 744061 (H2 O in HITRAN) lines in the database [83],
respectively. Also shown is the distribution of Planck’s emitted power Eλb at 5780
and 1800 K. The horizontal bars describe wavelength intervals that carry 99% of the
emitted power. Mean absorption coefficients of the gas mixture were calculated as
1
FORTRAN code used in this chapter has been programmed in the framework of: A. Camen-
zind, Spectral Radiation Properties of CO2 , CO and H2 O: Model Development and Implementa-
tion, Semester Thesis, ETH Zurich, 2004.
138 C. Radiation Absorption in the Gas Phase

5780 K H2 O
0
5
1800 K 10
eλb , W/(m2 · µm)
10

κ λ , 1/m
–5
10
0 10

CO

–5 –10
10 10
0.2 0.5 1 2 5 10 20
λ, µm

Figure C.1: Right axis: Absorption coefficients for CO and H2 O at 0.5 bar as a
function of the wavelength. Left axis: emissive power of a blackbody at 5780 K
and 1800 K as a function of the wavelength. 99 % of the total power is carried by
radiation in the intervals described by the horizontal bars. Solid and dotted lines
are for temperatures of 5780 K and 1800 K, respectively.

a sum of each specie’s contribution, while κλ was weighted by Eλb :


R∞
κλ Eλb dλ
κP = 0R ∞ (C.3)
0
Eλb dλ

Values of approx. 0.06 and 0.006 m−1 were obtained for radiation emitted at 1800
and 5780 K, respectively. Absorption by the gas is therefore neglected vis-à-vis
absorption by the particles.
List of Figures

2.1 Schematic representation of a refinery. Only units involved in the


conversion of crude oil to vacuum residue and delayed coke are shown. 11
2.2 Particle size distribution functions of the petcoke feedstock types
listed in Table 2.1. (a) shows the number density (also called popula-
tion density); (b) shows the volume density. Type 1: jet milled, type
2: ball milled, type 3: sieved with a 80 µm screen, type 4: sieved with
a 200 µm screen, type 5: as provided by PDVSA. . . . . . . . . . . . 14
2.3 SEM micrographs of petcoke samples used in Campaign 1 (feedstock
type 1). (a) and (b) show unreacted particles, (c) and (d) show par-
ticles after pyrolysis above 1400 K, (e) and (f) show particles after a
typical experimental run with combined pyrolysis and gasification and
carbon conversion XC =0.75. Magnification: (a),(c) and (e) 4’000x;
(b),(d) and (f) 10’000x . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Viscosity (a) and shear stress (b) as a function of the shear rate of a
water-coke slurry, measured for two types of feedstock [57]. The ar-
rows indicate the upward and downward curves, obtained by increas-
ing and subsequently decreasing the shear rate, respectively. Nicely
visible is the hysteresis loop, typical for thixotropic fluids. . . . . . . . 17

3.1 Scheme of the solar chemical reactor configuration used for the steam
gasification of carbonaceous materials at a power level of 5kW. The
location of the thermocouples (tc) of types K and S are indicated by
stars. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Feedstock injection setup used in Campaign 1. Petcoke particles are
fed by means of a brush conveyer and further entrained by the steam
injected from four tangential nozzles. . . . . . . . . . . . . . . . . . . 22
140 List of Figures

3.3 Feedstock injection setup used in Campaign 2. The coke-water slurry


is preheated in the tubes cast into the reactor’s insulation, the water
evaporates and the particles are injected into the cavity. . . . . . . . 23

3.4 Feedstock injection setup used in Campaign 3. The liquefied vacuum


residue is injected through a cold nozzle, the droplets are subsequently
heated, pyrolyzed and gasified. . . . . . . . . . . . . . . . . . . . . . . 24

3.5 Photograph of PSI’s high flux solar furnace. Main visible features are
the sun tracking heliostat, the Venetian-blind type shutter used to
control the incoming radiation and, inside the building, the parabolic
concentrator and the reactor. . . . . . . . . . . . . . . . . . . . . . . 25

3.6 Main components of the experimental setup at PSI’s high flux so-
lar furnace. Three different feeding systems are shown for the three
distinct experimental campaigns performed from 2004 to 2006. . . . . 26

4.1 Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 7 in Table 4.1). The dashed vertical
line indicates the start of the petcoke feeding, the two vertical dotted
lines delimit the interval considered for the steady state calculations. 34

4.2 Breakdown of the total input power into heating of the reactants
Q̇heating , chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad
and conduction losses Q̇cond for the 23 runs of Campaign 1. Also
plotted is the measured solar power input, Q̇solar with its respective
accuracy bounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.3 Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 10 in Table 4.2). The dashed verti-
cal line indicates the start of the slurry feeding, the two vertical dotted
lines delimit the interval considered for the steady state calculations. 38

4.4 Breakdown of the total input power into heating of the reactants
Q̇heating , chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad
and conduction losses Q̇cond for the 29 runs of Campaign 2. Also
plotted is the measured solar power input, Q̇solar with its respective
accuracy bounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
List of Figures 141

4.5 Carbon conversion and efficiencies for the 29 experimental runs of


Campaign 2 grouped by feedstock type and slurry molar ration H2 O/C.
Feedstock types 2, 3 and 4 correspond to initial mean particle diam-
eters of 6.7, 17.6 and 30.8 µm, respectively. . . . . . . . . . . . . . . . 39
4.6 Temperatures, solar power input and product gas composition for a
typical run of Campaign 1 (run # 12 in Table 4.3). The dashed
vertical line indicates the start of the petcoke feeding, the two ver-
tical dotted lines delimit the interval considered for the steady state
calculations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.7 Breakdown of the total input power into heating of the reactants
Q̇heating , chemical reaction enthalpy Q̇chem , reradiation losses Q̇rerad
and conduction losses Q̇cond for the 12 runs of Campaign 3. Also
plotted is the measured solar power input, Q̇solar with its respective
accuracy bounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.8 Fractions of hydrogen in the product gas derived by either pyrolysis
or gasification, and the feeding rates of reactants VR and steam for
experimental runs of Campaign 3. . . . . . . . . . . . . . . . . . . . . 44

5.1 Simulation setup used to determine the angular and radial distri-
butions of the incoming concentrated solar radiation at PSI’s solar
furnace [34]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.2 Solar radiative flux (left axis) and solar power (right axis) as a func-
tion of the aperture’s diameter for (a) PSI’s solar furnace, and (b)
CIEMAT’s solar tower. . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.3 Angular distribution of the concentrated solar radiation at the aper-
ture for PSI’s solar furnace and CIEMAT’s solar tower. . . . . . . . . 53
5.4 Lines of equal projected incident solar flux q̇solar in kW/m2 on planes
parallel to the aperture. Results are given for a 50 mm aperture and
a nominal power trough the aperture Q̇solar =5 kW. Also indicated is
the shape of the prototype reactor’s cavity (gray). . . . . . . . . . . . 54
5.5 Schematic of heat fluxes at the wall. Radiation emitted by particles
and other cavity wall elements, q̇ic , is incident from the inside (right),
while the wall itself emits radiation toward the inside. On the outside
(left) the reactor is cooled by natural convection. Finally, conduction
occurs through the three material layers of the wall. . . . . . . . . . . 55
142 List of Figures

5.6 Overall heat transfer coefficient U for conductive losses trough the
reactor walls as a function of the location along the reactor axis. Also
indicated is the longitudinal cross-section of the reactor’s cavity. . . . 56
5.7 Complex refractive index of quartz n + ik [72] and transmittance Tr w
for a 3 mm thick window and for three selected incident directions θ
as a function of wavelength λ. . . . . . . . . . . . . . . . . . . . . . . 57
5.8 Schematic of heat fluxes at the window. Concentrated solar radiation
q̇solar is incident from the outside of the reactor (left) whereas radiation
emitted by the cavity walls and particles, q̇ic , is coming from the inside
(right). The window itself emits radiation on both sides. Finally, the
window is cooled by forced convection on the outside. . . . . . . . . . 58
5.9 Volume density of the polydisperse particles used in Campaigns 1
(type 1) and 2 (type 3). Solid lines are measured values for the unre-
acted samples, dashed and dashed-dotted lines are calculated results
at carbon conversions of 0.5 and 0.75, respectively. . . . . . . . . . . . 59
5.10 Equivalent monodisperse diameters calculated with eqs (5.10) and
(5.26) of the polydisperse particles used in Campaigns 1 (type 1) and
2 (type 3) as a function of the carbon conversion. . . . . . . . . . . . 60
5.11 Spectral distribution of the absorption and scattering coefficients, cal-
culated for a typical volume fraction of 5·10−5 and carbon conversions
of 0.0, 0.5 and 0.75, for the feedstock used in (a) Campaign 1, and
(b) Campaign 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.12 Scattering phase function for the equivalent diameter D32 , for the
polydisperse medium (eq. (5.14)), and for the Henyey-Greenstein
approximation (eq. (5.15)), calculated for carbon conversions of (a)
0.0; and (b) 0.75, for the feedstock used in Campaign 1. Directions
θ = π/2 – π not shown in plot, because no backward scattering peak
was observed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6.1 Diagram of the considered process and it’s energy and mass flows used
in the simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.2 Carbon conversion (a), cavity (b) and shell (d) temperatures and
chemical efficiency (c) as a function of the carbon feeding rate and
the reactor volume for the 5 kW prototype reactor and the baseline
parameter listed in Table 6.1. . . . . . . . . . . . . . . . . . . . . . . 72
List of Figures 143

6.3 Predicted carbon conversion (a), cavity temperature (b), chemical


efficiency (c) and process efficiency (d) as a function of the carbon
feeding rate and the solar power for a scaled-up reactor. The baseline
parameter are listed in Table 6.1. . . . . . . . . . . . . . . . . . . . . 73

6.4 Numerically calculated (lumped parameters model) and experimental


measured data for the (a) 23 solar experimental runs of Campaign
1; and (b) 20 solar experimental runs of Campaign 2, ordered by
increasing Q̇solar . Shown are the average cavity and wall temperatures,
the carbon conversion and, the chemical and process efficiencies. Full
circles indicate numerically calculated values; open circles indicate the
experimentally measured data; the error bars indicate the propagated
inaccuracy of the input parameters. . . . . . . . . . . . . . . . . . . . 74

7.1 Axis-symmetric model domain, featuring five concentric cylindrical


compartments. Indicated are also the boundary and inlet conditions.
The grid spacing in x-direction is adapted to the expected tempera-
ture gradients, shown is an example for the Monte-Carlo solver. . . . 82

7.2 Typical finite volume considered in the simulation with faces ‘w’ and
‘e’, centerpoint ‘P’, finite length ∆x, and cross-section Sj . Also shown
are the respective face-centered and cell-centered variables. Advection
is solved in x-direction for each compartment j separately, whereas
radiative exchange is considered along both, the x and the r directions. 83

7.3 Flowchart of the Monte-Carlo raytracer used the calculate the radia-
tive source term in a domain composed of finite volumes. Note that
the volumes are delimited by either interfacial faces or/and domain
boundaries. Computation steps that involve one or several random
numbers are show by the bold boxes. . . . . . . . . . . . . . . . . . . 85

7.4 Typical volume element used in the MC solver. A ray emitted from a
boundary face (gray) with initial power qray and wavelength λ under-
going absorption and scattering is shown (dotted arrow). The par-
ticipating medium is described by κλ,g , κλ,s , and σsλ,g the absorption
coefficients for the gas and solid phase, and the scattering coefficient
of the solid phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
144 List of Figures

7.5 Schematic of a ray passing trough several elements of width Ln and ex-
tinction coefficient βn . Also shown is the residual penetration length
ln in the last element. . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.6 Flowchart of the coupled solver. The bold box describes the MC
raytracer described in Fig. 7.3. . . . . . . . . . . . . . . . . . . . . . . 92
7.7 Relative RMS error for the temperature and underrelaxation parame-
ter γ (left axis) and total emitted power, scaled by its maximal value,
used as an indicator for system convergence (right axis) as function
of the iteration number i. Values for run #12 of Campaign 1 with
ξ = 32. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.8 Relative grid convergence errors in radial direction, r , and in axial
direction, x , for carbon and steam conversions at the outlet as a
function of the grid refinement factor. The grid refinement factor 1,
2, 4 and 8 corresponds to 10, 20, 40, 80 elements in x-direction and
to 2, 4, 8 and 16 elements in r-direction. . . . . . . . . . . . . . . . . 95
7.9 Numerically calculated and experimental measured data for the (a)
23 solar experimental runs of Campaign 1; and (b) 20 solar experi-
mental runs of Campaign 2, ordered by increasing Q̇solar . Shown are
the average cavity and wall temperatures, and the carbon and steam
conversions at the outlet. Full circles indicate numerically calculated
values; open circles indicate the experimentally measured data; the
error bars indicate the propagated inaccuracy of the input parameters. 96
7.10 Numerically calculated temperatures profiles for the gaseous and solid
phases and the radiative source term along the reactor at two radial
positions: center (r = 0.0 m) and close to the wall (r = 0.025 m). The
baseline parameters listed in 7.1 have been employed for (a) Campaign
1; and (b) Campaign 2. . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7.11 Variation of the chemical composition (chemical species’ molar frac-
tions) along the reactor at two radial positions: center (r = 0.0 m)
and close to the wall (r = 0.025 m). The baseline parameters listed
in 7.1 have been employed for (a) Campaign 1; and (b) Campaign 2. . 100
7.12 2D contour map of the carbon conversion XC and reaction rate dXC /dt
(upper plot) and corresponding temperatures (lower plot). The base-
line parameters listed in 7.1 have been employed for (a) Campaign 1;
and (b) Campaign 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
List of Figures 145

8.1 Scheme of the up-scaled chemical reactor configuration for the solar
steam-gasification of coke, featuring a continuous gas-particle vortex
flow confined to a cavity receiver and directly exposed to concentrated
solar radiation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
8.2 (a) Temperature profiles (average values in the radial direction), and
(b) reaction extents along the reactor axis, for the baseline parame-
ters of Synpet5 (Table 8.1) and with monodisperse particles of initial
diameters 0.57, 0.96, 4.6, 62, and 176 µm, corresponding to curves 1,
2, 3, 4, and 5 respectively. . . . . . . . . . . . . . . . . . . . . . . . . 108
8.3 Reaction extent at the exit of the reactor as a function of the feed-
stock’s initial particle diameter for (a) Synpet5, and (b) Synpet300.
Curves are plotted for coke and water feeding rates corresponding to
0.5, 1, 2, and 4 times the baseline values of Table 8.1. Indicated are
the equivalent diameters for the polydisperse feedstock types 1, 3, and
5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
8.4 Reaction extent (left axis) and solar-to-chemical energy conversion
efficiency (right axis) as a function of the coke mass flow rate (nor-
malized to the baseline rate) for (a) Synpet5; and (b) Synpet300. The
feedstock types are 1, 3, and 5. The solid line is for optimal initial
particle diameters in the range 2-7 and 11-35 µm for Synpet5 and
Synpet300, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . 110
8.5 Solar-to-chemical energy conversion efficiency (a), and syngas yield
(b), as a function of input solar power for Synpet300 using feed-
stock type 5. The dotted, dash-dotted, and dashed lines represent
the performance for selected reaction extents of 0.8, 0.95 and 0.99,
respectively. The solid curve in both figures represents the locus of
maximum ηchem . Indicated are selected corresponding values of XC
in (a) and ṁcoke in (b). . . . . . . . . . . . . . . . . . . . . . . . . . . 111
8.6 Reactor cross-sections of constant volume where geometry ‘0’ corre-
sponds to the baseline dimensions (Table 8.1). . . . . . . . . . . . . . 112
8.7 Contour maps of reaction extent as a function of geometry (see Fig. 8.6)
and particle diameter for the baseline parameters listed in Table 8.1
and for: (a) Synpet5, and (b) Synpet300. . . . . . . . . . . . . . . . . 113

A.1 Scheme of the experimental setup. . . . . . . . . . . . . . . . . . . . . 124


146 List of Figures

A.2 Scheme of the suction thermocouple apparatus. . . . . . . . . . . . . 124


A.3 Left axis: experimentally determined and numerically computed Nus-
selt number as a function of Reynolds number. Right axis: corre-
sponding Graetz numbers, normalized by Gzlim . . . . . . . . . . . . . 129
A.4 Experimental data (solid lines) and calculated gas temperatures (dashed
lines) as a function of the suction mass flow rate at a nominal furnace
temperature of 1223 and N2 mass flow rate of 10 ln /min. . . . . . . . 131
A.5 Relative differences between the gas temperature simulated by CFD
(TgasCFD ) and the bare thermocouple reading (Tbare ), the shielded ther-
mocouple reading (Ttc ), and calculated gas temperature (Tgas ), at dif-
ferent Reynolds numbers. See definitions of δbare , δtc , and δmethod in
equations (A.9), (A.10), and (A.11), respectively. . . . . . . . . . . . 132

C.1 Right axis: Absorption coefficients for CO and H2 O at 0.5 bar as a


function of the wavelength. Left axis: emissive power of a blackbody
at 5780 K and 1800 K as a function of the wavelength. 99 % of the
total power is carried by radiation in the intervals described by the
horizontal bars. Solid and dotted lines are for temperatures of 5780
K and 1800 K, respectively. . . . . . . . . . . . . . . . . . . . . . . . 138
List of Tables

2.1 Coke feedstock types used in the experiments. . . . . . . . . . . . . . 13


2.2 Approximate main elemental chemical composition (ultimate analy-
sis), low heating value and molar ratios H/C and O/C for PD coke. . 16
2.3 Approximate main elemental chemical composition (ultimate analy-
sis), low heating value and molar ratios H/C and O/C for vacuum
residue. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.1 Theoretical maximal solar-to-chemical energy conversion efficiency


ηchem for coke and vacuum residue for the 5 kW prototype reactor
and its 300 kW scale-up as a function of the cavity temperature. . . . 30

4.1 Steam and petcoke feeding rates, composition of the product gas,
nominal temperatures and performance parameters for 23 valid runs
of experimental campaign 1 conducted with dry petcoke particles. . . 36
4.2 Water and petcoke feeding rates, composition of the product gas,
nominal temperatures and performance parameters for 29 valid runs
of experimental campaign 2 conducted with a coke-water slurry. . . . 41
4.3 Steam and VR feeding rates, composition of the product gas, nom-
inal temperatures and performance parameters for 12 valid runs of
experimental campaign 3 conducted with liquefied VR. . . . . . . . . 46
4.4 Average operational parameters and results of the solar experimental
campaigns conducted with dry coke powder, cokewater slurry, and VR. 47

5.1 Complex refractive index of coke [14, 29]. . . . . . . . . . . . . . . . . 61


5.2 Arrhenius parameters of the kinetic rate constants for the steam gasi-
fication of coke [100]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
148 List of Tables

6.1 Baseline model parameters for the studies conducted for the 5 kW
lab scale reactor and it’s scale-up. . . . . . . . . . . . . . . . . . . . . 71

7.1 Baseline model parameters for two representative solar experimental


runs for Campaigns 1 and 2. . . . . . . . . . . . . . . . . . . . . . . . 97
7.2 Relative sensitivity of input/output parameters for both experimental
campaigns. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

8.1 Baseline operational parameters for the two reactors analyzed. . . . . 106

A.2 Experimental operating conditions and measurement results for the


calibration of the suction thermocouple apparatus. . . . . . . . . . . . 127
A.3 Geometrical dimensions of the suction thermocouple apparatus and
material properties for air, N2 , and Mo. . . . . . . . . . . . . . . . . . 128
A.4 Parameters for the empirical correlations of the Nusselt number (c1
and c2 in eq. (A.2)), and of the effective thermal conductivity (c3 and
c4 in eq. (A.6), and regression accuracy R2 . . . . . . . . . . . . . . . . 128
A.5 Experimental operating conditions, measurement results, and calcu-
lated gas temperatures with N2 at different furnace temperatures. . . 130
A.6 Measured temperatures of the working tube in the furnace as a func-
tion of the axial position x, starting from the inlet. The center of the
furnace as well as the entrance of the suction thermocouple apparatus
is at position x = 0.195 m. . . . . . . . . . . . . . . . . . . . . . . . . 132

B.1 Influence of the inaccurate positioning of the aperture with respect


to the focus of the parabolic concentrator (spot). . . . . . . . . . . . 135
Bibliography

[1] S. Abanades and G. Flamant. Solar hydrogen production from the thermal
splitting of methane in a high temperature solar chemical reactor. Solar En-
ergy, 80:1321–1332, 2006. 1.3

[2] R. Alvarez. An experimental investigation of the steam gasification of petcoke.


Semester Project, ETH Zurich, 2003. 3.1

[3] R. Alvarez. Solar thermal steam gasification of petcoke in an opaque fluidized


bed reactor. Master’s thesis, ETH Zurich, 2004. 3.1

[4] M. J. Antal, L. Hofmann, and J. R. Moreira. Design and operation of a solar


fired biomass flash pyrolysis reactor. Solar Energy, 30:299–312, 1983. 1.3

[5] J. Ballestrı́n and R. Monterreal. Hybrid heat flux measurement system for
solar central receiver evaluation. Energy, 29:915–924, 2004. B

[6] I. K. Basily and K. Thamer. Pyrolysis products dependence on the nature of


heavy hydrocarbon feedstock. Applied Catalysis, 73:125–133, 1991. 4.3

[7] W. H. Beattie, R. Berjoan, and J.-P. Coutures. High temperature solar pyrol-
ysis of coal. Solar Energy, 31:137–143, 1983. 1.3

[8] R. Berber and E. A. Fletcher. Extracting oil from shale using solar energy.
Energy, 13:12–23, 1988. 1.3

[9] L. G. Blevins and W. M. Pitts. Modeling of bare and aspirated thermocouples


in compartment fires. Fire Safety Journal, 33:239–259, 1999. A.2

[10] F. Bögner and K. Wintrup. Handbook of Synfuels Technology, chapter The


Fluidized-Bed Coal Gasification Process (Winkler Type), pages 111–125.
McGraw-Hill, New York, 1984. 1.2
150 Bibliography

[11] C. F. Bohren and D.R.Huffman. Absorption and Scattering of Light by Small


particles. Wiley, New York, 1983. 5.2.1, 5.2.1

[12] A. W. Bowman and A. Azzalini. Applied Smoothing Techniques for Data


Analysis The Kernel Approach with S-Plus Illustrations. Clarendon Press,
1997. 7.3.3

[13] R. Buck, J. F. Muir, R. E. Hogan, and R. D. Skocypec. Carbon dioxide


reforming of methane in a solar volumetric receiver/reactor: the CAESAR
project. Solar Energy Materials, 24:449–463, 1991. 1.3

[14] L. D. Burak. Radiative properties of coke particles of coal-dust flame.


Inzhenerno-Fizicheskii Zhurnal, 45:297–302, 1983. 5.2.1, 5.1, C

[15] V. L. Cabal. Climate change and European cleaner coal technology. Energy
& Environment, 14:3–15, 2003. 1

[16] P. Charvin, S. Abanades, G. Flamant, and F. Lemort. Two-step water split-


ting thermochemical cycle based on iron oxide redox pair for solar hydrogen
production. Energy, 32:1124–1133, 2007. 1.3

[17] P. R. N. Childs, J. R. Greenwood, and C. A. Long. Review of temperature


measurement. Review of Scientific Instruments, 71:2959–2978, 2000. A.2

[18] J. K. Dahl, K. J. Buechler, R. Finley, T. Stanislaus, A. W. Weimer,


A. Lewandowski, C. Bingham, A. Smeets, and A. Schneider. Rapid solar-
thermal dissociation of natural gas in an aerosol flow reactor. Energy, 29:715–
725, 2004. 1.3

[19] J. K. Dahl, A. W. Weimer, A. Lewandowski, C. Bingham, F. Bruetsch, and


A. Steinfeld. Dry reforming of methane using a solar-thermal aerosol flow
reactor. Industrial & Engineering Chemistry Research, 43:5489–5495, 2004.
1.3

[20] H. Dinçer, F. Boylu, A. Sirkeci, and G. Ateşok. The effect of chemicals on the
viscosity and stability of coal water slurries. International Journal of Mineral
Processing, 70:41–51, 2003. 2.1.1

[21] M. Dry. The SASOL route to fuels. ChemTech, 12:744–750, 1982. 1.2
Bibliography 151

[22] R. Edinger and S. Kaul. Humankind’s detour toward sustainability: past,


present, and future of renewable energies and electric power generation. Re-
newable and Sustainable Energy Reviews, 4:295–313, 2000. 1

[23] F. O. Ernst, A. Tricoli, , S. E. Pratsinis, and A. Steinfeld. Co-synthesis of H2


and ZnO by in-situ Zn aerosol formation and hydrolysis. AIChE, 52:3297–3303,
2006. 1.3

[24] J. T. Farmer and J. R. Howell. Advances in heat transfer, chapter Comparison


of Monte Carlo Strategies for Radiative Transfer in Participating Media, pages
333–429. Academic Press, 1998. 7.3.1

[25] M. Fasciana and F. Noembrini. Solar pet-coke gasification, direct irradiation.


Master’s thesis, ETH Zurich, 2003. 3.1

[26] M. Flechsenhar and C. Sasse. Solar gasification of biomass using oil shale and
coal as candidate materials. Energy, 20:803–810, 1995. 1.3

[27] E. Fletcher and R. Moen. Hydrogen and oxygen from water. the use of solar
energy in a one-step effusional process is considered. Science, 197:1050–1056,
1977. 1.3

[28] E. A. Fletcher. Solarthermal processing: A review. Journal of Solar Energy


Engineering, 123:63–74, 2001. 1.3

[29] F. Goodarzi. Optical properties of vitrinite carbonized at different pressures.


Fuel, 64:156–162, 1985. 5.2.1, 5.1, C

[30] M. R. Gray. Upgrading Petroleum Residues and Heavy Oils. Taylor & Francis,
CRC Press, 1994. 4.3

[31] D. W. Gregg, R. W. Taylor, J. H. Campbell, J. R. Taylor, and A. Cotton. Solar


gasification of coal, activated carbon, coke and coal and biomass mixtures.
Solar Energy, 25:353–364, 1980. 1.3

[32] Y. Guo, C. Chan, and K. Lau. Numerical studies of pulverized coal combustion
in a tubular coal combustor with slanted oxygen jet. Fuel, 82:893–907, 2003.
A.2
152 Bibliography

[33] R. P. Gupta, T. F. Wall, and J. S. Truelove. Radiative scatter by fly ash


in pulverized-coal-fired furnaces: Application of the Monte Carlo method to
anisotropic scatter. International Journal of Heat and Mass Transfer, 26:1649–
1660, 1983. 7.3.1

[34] P. Haueter, T. Seitz, and A. Steinfeld. A new high-flux solar furnace for high-
temperature thermochemical research. Journal of Solar Energy Engineering,
121:77–80, 1999. 3.3, 5.1, C

[35] M. Haug and C. Difiglio. Biofuels for Transport - An International Perspective.


Internation Energy Agency, 2004. 1

[36] L. G. Henyey and J. L. Greenstein. Diffuse radiation in the galaxy. Astrophys-


ical Journal, 88:70–83, 1940. 5.2.1

[37] C. Higman and M. van der Burg. Gasification. Gulf Professional Publishing,
2003. 1.2

[38] S. Hiller, O. Deussen, and A. Keller. Tiled blue noise samples. In Vision,
Modeling, and Visualization 2001, 2001. 7.3.3

[39] A. W. D. Hills and A. Paulin. The construction and calibration of an inex-


pensive microsuction pyrometer. Journal of Physics E, 2:713–717, 1969. A.2

[40] D. Hirsch and A. Steinfeld. Radiative transfer in a solar chemical reactor


for the co-production of hydrogen and carbon by thermal decomposition of
methane. Chemical Engineering Science, 59:5771–5778, 2004. 7.3.1, A.2

[41] D. Hirsch and A. Steinfeld. Solar hydrogen production by thermal decom-


position of natural gas using a vortex-flow reactor. International Journal of
Hydrogen Energy, 29:47–55, 2004. 1.3

[42] D. Hirsch, P. von Zedtwitz, T. Osinga, J. Kinamore, and A. Steinfeld. A new


75 kW high-flux solar simulator for high-temperature thermal and thermo-
chemical research. Journal of Solar Energy Engineering, 125:117–131, 2003.
3.1

[43] J. R. Howell. Thermal radiation in participating media, the past, the present
and some possible futures. Journal of Heat Transfer, 1100:1220–1229, 1988.
7.3.1
Bibliography 153

[44] F. P. Incropera and D. P. DeWitt. Fundamentals of Heat and Mass Transfer.


John Wiley & Sons, 2002. 5.1.3, 7.1, 7.1, 7.2, 7.2, A.4

[45] E. Jochem, G. Andersson, D. Favrat, H. Gutscher, K. Hungerbhler, P. R. von


Rohr, D. Spreng, A. Wokaun, and M. Zimmermann. Steps towards a sus-
tainable development. a white book for R&D of energy-efficient-technologies.
Technical report, CEPE/ETH Zurich and Novatlantis, Zurich, 2004. 1

[46] T. B. Johansson and L. Burnham. Renewable Energy: Sources for Fuels and
Electricity. Island Press, 1993. 1

[47] R. Kandiyoti, A. Herod, and K. Bartle. Solid Fuels and Heavy Hydrocarbon
Liquids - Thermal Characterization and Analysis. Elsevier, 2006. 1

[48] T. Kappauf and E. A. Fletcher. Hydrogen and sulfur from hydrogen sulfide-VI.
solar thermolysis. Energy, 14:443–449, 1989. 1.3

[49] J. Katzer. The Future of Coal - an Interdisciplinary MIT Study. Massachusetts


Institute of Technology, 2007. 1

[50] M. B. Khalil, F. M. El-Mahallawy, and S. A. Fareg. Accuracy of temperature


measurements in furnaces. Letters in Heat and Mass Transfer, 3:421–432,
1976. A.2

[51] S. Kim and B. E. Dale. Global potential bioethanol production from wasted
crops and crop residues. Biomass and Bioenergy, 26:361–375, 2004. 1

[52] T. Kodama. High-temperature solar chemistry for converting solar heat to


chemical fuels. Progress in Energy and Combustion Science, 29:567–597, 2003.
1.3

[53] T. Kodama, A. Kiyama, T. Moriyama, and O. Mizuno. Solar methane reform-


ing using a new type of catalytically-activated metallic foam absorber. Journal
of Solar Energy Engineering, 126:808–811, 2004. 1.3

[54] A. Kogan. Direct solar thermal splitting of water and on-site separation of the
products. International Journal of Hydrogen Energy, 23:89–98, 1998. 1.3

[55] A. K. Kolar and R. Sundaresan. Heat transfer characteristics at an axial tube


in a circulating fluidized bed riser. International Journal of Thermal Sciences,
41:673–681, 2002. A.2
154 Bibliography

[56] T. Kollig and A. Keller. Efficient multidimensional sampling. In EURO-


GRAPHICS 2002, volume 21, 2002. 7.3.3

[57] F. Kritter. Injection system of coal water slurries for solar reactors. Master’s
thesis, ETH Zurich, 2005. 2.4, C

[58] S. Lee. Alternative Fuels. Taylor & Francis, 1996. 1, 1.2

[59] O. Levenspiel. Chemical Reaction Engineering. Wiley, New York, 1999. 3.4.4

[60] W. Lipiński, A. Z’Graggen, and A. Steinfeld. Transient radiation heat transfer


within a non-gray non-isothermal absorbing-emitting-scattering suspension of
reacting particles undergoing shrinking. Numerical Heat Transfer, Part B,
47:443–457, 2005. 7.3.1, A.2

[61] F. C. Lockwood, A. P. Salooja, and S. A. Syed. A prediction method for


coal-fired furnaces. Combustion and Flame, 38:1–15, 1980. A.2

[62] M. Lundberg. Model calculations on some feasible two-step water splitting


processes. International Journal of Hydrogen Energy, 18:369–376, 1993. 1.3

[63] M. Luo. Effects of radiation on temperature measurement in a fire environ-


ment. Fire Sciences, 15:443–461, 1997. A.2

[64] J. Marakis, C. Papapavlou, and E. Kakaras. A parametric study of radiative


heat transfer in pulverized coal furnaces. International Journal of Heat and
Mass Transfer, 43:2961–2971, 2000. A.2

[65] J. Matsunami, S. Yoshida, O. Yokota, M. Nezuka, M. Tsuji, and Y. Tamaura.


Gasification of waste tyre and plastic (pet) by solar thermochemical process
for solar energy utilization. Solar Energy, 65:21–23, 1999. 1.3

[66] L. Michalski, K. Eckersdorf, J. Kucharski, and J. McGhee. Temperature Mea-


surement, chapter 17. Temperature Measurement of Fluids, pages 361–380.
John Wiley & Sons Ltd, 2001. A.2

[67] D. P. Mitchell. Generating antialiased images at low sampling densities. Com-


puter Graphics, 21:65–72, 1987. 7.3.3

[68] M. F. Modest. Radiative heat transfer. Academic Press, 2003. 3.1, 5.1.3, 5.2.1,
5.2.1
Bibliography 155

[69] J. Nelder and R.Mead. A simplex method for function minimization. Computer
Journal, 7:308–313, 1965. 6.1, A.4

[70] T. Osinga, U. Frommherz, A. Steinfeld, and C. Wieckert. Experimental inves-


tigation of the solar carbothermic reduction of ZnO using a two-cavity solar
reactor. Journal of Solar Energy Engineering, 126:633–636, 2004. 1.3

[71] T. Osinga, G. Olalde, and A. Steinfeld. Solar carbothermal reduction of ZnO:


Shrinking packed-bed reactor modeling and experimental validation. Industrial
& Engineering Chemistry Research, 43:7981–7988, 2004. A.2

[72] E. D. Palik. Handbook of optical constants of solids. Academic Press, 1985.


5.1.3, 5.7, C

[73] H. Y. Park and T. H. Kim. Non-isothermal pyrolysis of vacuum residue (VR) in


a thermogravimetric analyzer. Energy Conversion and Management, 47:2118–
2127, 2006. 4.3

[74] S. V. Patankar. Numerical heat transfer and fluid flow. Hemisphere Publishing
Corporation; New York, 1980. 7.3.2

[75] International Energy Agency. Industry Attitudes to Combined Cycle Clean


Coal Technologies. 1994. 1

[76] Lurgi — Kohle- und Mineralöltechnik. Lurgi Handbuch. Lurgi


Gesellschaften,Frankfurt a.M., 1970. 1.2

[77] Plataforma Solar de Almerı́a. Annual Report 2006. CIEMAT, Spain, 2006.
8.1

[78] World Energy Council. 2007 Survey of Energy Resources. Regency House,
London, 2007. 1

[79] J. Rezaiyan and N. P. Cheremisinoff. Gasification Technologies. Taylor &


Francis, 2005. 1.2, 2.1, 4.1

[80] P. J. Roache. Verification and Validation in Computational Science and En-


gineering. Hermosa Publishers, Albuquerque, 1998. 7.3.3
156 Bibliography

[81] M. Roeb, A. Noglik, N. Monnerie, M. Schmitz, C. Sattler, G. Cerri,


G. de Maria, A. Giovanelli, A. Orden, D. de Lorenzo, J. Cedillio, A. le Digou,
and J.-M. Borgard. Development and verification of process concepts for the
splitting of sulphuric acid by concentrated solar radiation. In M. Romero, edi-
tor, 13th International Symposium on Concentrated Solar Power and Chemical
Energy Technologies, 2006. 1.3

[82] H.-H. Rogner. An assessment of world hydrocarbon resources. Annual Review


of Energy and the Environment, 22:217–262, 1997. 1

[83] L. S. Rothman, C. P. Rinsland, A. Goldman, S. T. Massie, D. P. Edwards,


J.-M. Flaud, A. Perrin, C. Camy-Peyret, V. Dana, Y.-Y. Mandin, J. W.
Schroeder, R. R. Gamache, R. B. Wattson, K. Yoshino, K. V. Chance, K. W.
Jucks, L. R. Brown, V. Nemtchinovi, and P. Varanasi. The HITRAN molecu-
lar spectroscopic database and HAWKS (HITRAN atmospheric workstation):
1996 edition. Journal of Quantitative Spectroscopy & Radiative Transfer,
60:665–710, 1998. 7.2, C, C

[84] D. Ruppert and M. P. Wand. Multivariate locally weighted least squares


regression. The Annals of Statistics, 22:1346–1370, 1994. 7.3.3

[85] P. L. Russell. Oil shales of the world, their origin, occurrence and exploitation.
Pergamon Press, 1990. 1

[86] A. Saltelli, S. Tarantola, F. Campolongo, and M. Ratto. Sensitivity analysis


in practice: a guide to assessing scientific models. John Wiley & Sons Ltd,
2004. 7.4

[87] W. G. Schlinger. Handbook of Synfuels Technology, chapter The Texaco Coal


Gasification Process. McGraw-Hill, New York, 1984. 1.2

[88] E. Schlünder, editor. VDI-Wärmeatlas, Berechnungsblätter für den


Wärmeübergang. VDI-Verlag, Düsseldorf, 1984. 7.2

[89] L. Schunk, P. Haeberling, S. Wepf, D. Wuillemin, A. Meier, and A. Steinfeld.


A solar receiver-reactor for the thermal dissociation of zinc oxide. Journal of
Solar Energy Engineering, 2008. 1.3
Bibliography 157

[90] P. F. Sens and J. K. Wilkinson. Coal-water mixtures. Elsevier, London, 1989.


2.1.1

[91] R. Siegel and J. R. Howell. Thermal radiation heat transfer. Taylor and
Francis, 2002. 5.2.1, 7.3.1, 7.3.1

[92] R. Speck. Solar petcoke gasification, indirect irradiation. Master’s thesis, ETH
Zurich, 2003. 3.1

[93] A. Steinfeld. Solar thermochemical production of hydrogen - a review. Solar


Energy, 78:603–615, 2005. 1

[94] A. Steinfeld and R. Palumbo. Encyclopedia of Physical Science & Technology,


Vol. 15, chapter Solar Thermochemical Process Technology, pages 237–256.
Academic Press, 2001. 1.3

[95] R. W. Taylor, R. Berjoan, and J. P. Coutures. Solar gasification of carbona-


ceous materials. Solar Energy, 30:513–525, 1983. 1.3

[96] C. L. Tien. Thermal radiation in packed and fluidized beds. Journal of Heat
transfer, 110:1230–1242, 1988. 7.3.1

[97] M. Toda, M. Kuriyama, H. Konno, and T. Honma. The influence of parti-


cle size distribution of coal on the fluidity of coal-water mixtures. Powder
Technology, 55:241 – 245, 1988. 2.1.1

[98] Y. Touloukian, R. Powell, C. Ho, and P. Klemens. Thermophysical Properties


of Matter, Volume 1. Thermal Conductivity: Metallic Elements and Alloys.
Plenum Press, New York, 1970. A.3

[99] Y. S. Touloukian and C. Y. Ho. Thermophysical Properties of Matter, Volume


7. Thermal Radiative Properties: Metallic Elements and Alloys. Plenum Press,
New York, 1970. A.3

[100] D. Trommer. Thermodynamic and Kinetic Analyses of Solar Thermal Gasifi-


cation of Petroleum Coke. PhD thesis, ETH Zurich, 2006. 1.1, 5.2, 5.3, 5.3, 9,
C

[101] D. Trommer, F. Noembrini, M. Fasciana, D. Rodriguez, A. Morales,


M. Romero, and A. Steinfeld. Hydrogen production by steam-gasification of
158 Bibliography

petroleum coke using concentrated solar power - I. thermodynamic and kinetic


analyses. International Journal of Hydrogen Energy, 30:605–618, 2005. 1, 1.3,
4.1, 4.2, 4.3, 7.4

[102] D. Trommer and A. Steinfeld. Kinetic modeling for the combined pyrolysis
and steam-gasification of petroleum coke and experimental determination of
the rate constants by dynamic thermogravimetry in the 500-1520 K range.
Energy & Fuels, 20:1250–1258, 2006. 1.1, 4.2, 5, 5.3, 5.3

[103] H. R. Tschudi and G. Morian. Pyrometric temperature measurements in solar


furnaces. Journal of Solar Energy Engineering, 123:164–170, 2001. 3.1

[104] S. Ulmer, E. Lüpfert, M. Pfänder, and R. Buck. Calibration corrections of


solar tower flux density measurements. Energy, 29:925–933, 2004. B

[105] H. Usui, T. Saeki, K. Hayashi, and T. Tamura. Sedimentation stability and


rheology of coal water slurries. International Journal of Coal Preparation and
Utilization, 18:201–214, 1997. 2.1.1

[106] N. B. Vargaftik. Tables on the Thermophysical Properties of Liquids and


Gases. Hemisphere Publishing Corporation, 1975. A.3, A.7

[107] I. Vishnevetsky and M. Epstein. Production of hydrogen from solar zinc in


steam atmosphere. International Journal of Hydrogen Energy, 32:2791–2802,
2007. 1.3

[108] R. Viskanta and M. P. Mengüc. Radiation heat transfer in combustion systems.


Progress in Energy and Combustion Science, 13:97–160, 1987. A.2

[109] E. V. Vogt, M. J. V. D. Burgt, J. H. Chesters, and J. C. Hoogendoorn. De-


velopment of the Shell-Koppers coal gasification process. Philosophical Trans-
actions of the Royal Society of London. Series A, Mathematical and Physical
Sciences, 300:111–120, 1981. 1.2

[110] P. von Zedtwitz, W. Lipinski, and A. Steinfeld. Numerical and experimental


study of gas-particle radiative heat exchange in a fluidized-bed reactor for
steam-gasification of coal. Chemical Engineering Science, 62:599–607, 2007.
7.3.1, 7.4
Bibliography 159

[111] P. von Zedtwitz, J. Petrasch, D. Trommer, and A. Steinfeld. Hydrogen pro-


duction via the solar thermal decarbonization of fossil fuels. Solar Energy,
80:1333–1337, 2006. 1.3

[112] P. von Zedtwitz and A. Steinfeld. The solar thermal gasification of coal - energy
conversion efficiency and CO2 mitigation potential. Energy, 28:441–456, 2003.
1, 1.3

[113] D. V. Walters and R. O. Buckius. Monte Carlo methods for radiative heat
transfer in scattering media. Annual Review of Heat Transfer, 5:131–176, 1992.
7.3.1, 7.3.1

[114] W. T. Welford and R. Winston. High collection nonimaging optics. Academic


Press, New York, 1989. 4.1

[115] C. Wieckert, U. Frommherz, S. Kräupl, E. Guillot, G. Olalde, M. Epstein,


S. Santén, T. Osinga, and A. Steinfeld. A 300 kW solar chemical pilot plant
for the carbothermic production of zinc. Journal of Solar Energy Engineering,
129:190–196, 2007. 1.3

[116] M. S. Wojtan and K. A. G. Jones. Improvements in the design and operation


of a suction pyrometer. Measurement and Control, 14:301–305, 1981. A.2

[117] W. D. Wood, H. W. Deem, and C. Lucks. Thermal Radiative Properties.


Plenum Press, New York, 1964. A.3

[118] D. Xie, B. Bowen, J. Grace, and C. Lim. Two-dimensional model of heat trans-
fer in circulating fluidized beds. Part II: Heat transfer in a high density CFB
and sensitivity analysis. International Journal of Heat and Mass Transfer,
46:2193–2205, 2003. A.2

[119] Y. S. Yang, J. R. Howell, and D. E. Klein. Radiative heat transfer through a


randomly packed bed of spheres by the Monte Carlo method. Journal of Heat
Transfer, 105:325–332, 1983. 7.3.1

[120] A. Z’Graggen, H. Friess, and A. Steinfeld. Gas temperature measurement


in thermal radiating environments using a suction thermocouple apparatus.
Measurement Science and Technology, 18:3329–3334, 2007. 1.1
160 Bibliography

[121] A. Z’Graggen, P. Haueter, G. Maag, M. Romero, and A. Steinfeld. Hydrogen


production by steam-gasification of carbonaceous materials using concentrated
solar energy - IV. reactor experimentation with vacuum residue. International
Journal of Hydrogen Energy, 33:679–684, 2008. 1.1, 1.3

[122] A. Z’Graggen, P. Haueter, G. Maag, A. Vidal, M. Romero, and A. Steinfeld.


Hydrogen production by steam-gasification of petroleum coke using concen-
trated solar power — III. reactor experimentation with slurry feeding. Inter-
national Journal of Hydrogen Energy, 32:992–996, 2007. 1.1, 1.3

[123] A. Z’Graggen, P. Haueter, D. Trommer, M. Romero, J. C. de Jesus, and


A. Steinfeld. Hydrogen production by steam-gasification of petroleum coke
using concentrated solar power — II. reactor design, testing, and modeling.
International Journal of Hydrogen Energy, 31:797–811, 2006. 1.1, 1.3, 6.1

[124] A. Z’Graggen and A. Steinfeld. Radiative exchange within a two-cavity con-


figuration with a spectrally selective window. Journal of Solar Energy Engi-
neering, 126:819–822, 2004. 4.2, 4.3

[125] A. Z’Graggen and A. Steinfeld. Heat and mass transfer analysis of a suspension
of reacting particles subjected to concentrated solar radiation — application
to the steam-gasification of carbonaceous materials. International Journal of
Heat and Mass Transfer, 2007, submitted. 1.1

[126] A. Z’Graggen and A. Steinfeld. Hydrogen production by steam-gasification


of carbonaceous materials using concentrated solar energy — V. reactor mod-
eling, optimization, and scale-up. International Journal of Hydrogen Energy,
2008, submitted. 1.1
Curriculum Vitae

Name: Andreas Z’Graggen


Nationality: Swiss
Citizen of: Schattdorf (UR)
Date of birth: July 25, 1978

2003-2008 Doctoral studies at the Professorship in Renewable En-


ergy Carriers, ETH Zurich; supervision: Prof. Dr. Aldo
Steinfeld
1997-2003 Diploma studies in Mechanical Engineering at ETH Zurich;
majors in Energy Technology and Product Development
1993-1997 Maturità tipo C (scientifico), Liceo Cantonale, Lugano

You might also like