Sei sulla pagina 1di 6

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/266135729

Predictive pyrolysis process modelling In Aspen


Plus
CONFERENCE PAPER JANUARY 2014

CITATIONS

READS

428

3 AUTHORS:
Jens F. Peters

Diego Iribarren

Karlsruhe Institute of Technology

Madrid Institute for Advanced Studies

11 PUBLICATIONS 50 CITATIONS

43 PUBLICATIONS 527 CITATIONS

SEE PROFILE

SEE PROFILE

Javier Dufour
King Juan Carlos University
80 PUBLICATIONS 510 CITATIONS
SEE PROFILE

Available from: Jens F. Peters


Retrieved on: 20 October 2015

PREDICTIVE PYROLYSIS PROCESS MODELLING IN ASPEN PLUS


Jens F. Peters1,*, Diego Iribarren1 and Javier Dufour1,2
2

1
Systems Analysis Unit. Instituto IMDEA Energa. Mstoles 28935 (Spain).
Department of Chemical and Energy Technology. Rey Juan Carlos University. Mstoles 28933 (Spain).
* Corresponding author: Tel.: +34-91 737 11 46; E-mail address: jens.peters@imdea.org

ABSTRACT: This work presents a novel Aspen Plus model of pyrolysis processes for lignocellulosic feedstocks.
Based on kinetic reaction mechanisms, the simulation calculates product yields and composition depending on reactor
conditions (temperature, residence time, flue gas flow rate) and feedstock composition (cellulose, hemicellulose and
lignin fraction, atomic composition, ash and alkali metal content). The produced bio-oil is modelled with a high level
of detail (33 compounds including organic acids, aldehydes, alcohols, ketenes, phenols, sugar derivatives and
degraded lignin), and the char product shows realistic atomic compositions. N, S and Cl trace element release is taken
into account and the corresponding emissions caused by the process can be determined.
Numerous simulation runs are made in order to cross-check the simulation results with experimental data based on
published literature. The results show a high correlation of the results for the most common pyrolysis processes (e.g.,
fast pyrolysis in bubbling or circulating fluidised beds), which is somewhat decreasing for slow pyrolysis processes.
The simulation model is found to be suitable for predicting pyrolysis yields and products within the typical range of
operation for pyrolysis processes.
KEYWORDS: biomass, fast pyrolysis, lignocellulosic sources, process simulation, pyrolysis oil, reactor modelling

INTRODUCTION

Second-generation biofuels are seen as a solution for


further increasing the share of renewable energies in the
transport sector while reducing the negative impacts
associated with conventional biofuels [1]. In fact, there is
an important potential of lignocellulosic biomass from
forest residue, agricultural waste and energy crops which
is still unused and potentially suitable for bioenergy
production with low environmental impact [2].
Maximizing the use of residual biomass is necessary for
fulfilling the targets regarding biofuel share and
greenhouse gas emission reduction set up in the proposal
for the new renewable energy directive [1]. Nevertheless,
converting lignocellulosic biomass into liquid fuels is not
an easy task and adequate technologies are needed.
Pyrolysis, the thermal decomposition under intermediate
temperature and in oxygen-free atmosphere, is one of the
options for this purpose. Pyrolysis of lignocellulosic
biomass yields gases, liquids (the so-called bio-oil) and a
carbonaceous residue, char. The obtained yields depend
on feedstock composition and pyrolysis conditions, with
a typical maximum bio-oil yield of around 70 wt% under
fast pyrolysis conditions around 500C and residence
times of 1s [3,4]. Nevertheless, the obtained bio-oil is of
poor quality and requires significant improvement before
it can be used as transport fuel [46].
2

BACKGROUND

Pyrolysis is a possible pathway for converting


biomass into a liquid. Nevertheless, biomass pyrolysis is
mainly in the research stage and almost no commercial
pyrolysis installations exist yet [7,8]. Furthermore, as
bio-oil is a very complex liquid made up of hundreds of
compounds, hardly any exhaustive studies with
information about the detailed compositions of bio-oils
obtained from different types of feedstock are available.
Technical and environmental assessments therefore
mostly use a simple top-down approach to adjust the
pyrolysis products of their model to existing literature
data. However, pyrolysis of different types of biomass
(woody, herbaceous) gives different yields and product

compositions, why a simple model is not sufficient for a


detailed assessment of different processes dealing with
varying types of feedstocks. As system analysis for
research-stage processes is often based on process
simulation, there is a need for a flexible simulation that
facilitates the use of different new materials.
3

ASPEN PLUS SIMULATION

This work presents a novel model of a pyrolysis


reactor based on a kinetic reaction approach implemented
in Aspen Plus. Unlike other works, which implement
the pyrolysis reactor as a black box unit giving an a
priori defined pyrolysis product composition, the use of a
kinetic approach permits a really predictive simulation,
which estimates yield and composition of the pyrolysis
products depending on reactor conditions and feedstock
composition without the need of experimental data as
input. It requires the fractional composition of the
biomass (cellulose, hemicellulose, lignin) and its atomic
composition plus its ash and alkali metal content, and
calculates the pyrolysis products based on a superseded
reaction model for the cellulose, hemicellulose and lignin
fractions of the biomass. The modelled bio-oil is of high
complexity and contains 33 compounds including organic
acids, aldehydes, alcohols, ketenes, phenols, sugar
derivatives and degraded lignin, thus giving a good
approximation of the real bio-oil composition. Trace
element release is taken into account. Hence, bio-oil, gas
and char show a realistic O, N, S and Cl content, which is
of importance for determining the emissions caused by
the process.
The pyrolysis reactor consists of three subsequent
reactors, which are necessary to model all the reactions
occurring during biomass pyrolysis with the required
level of detail. The Aspen Plus flowsheet of the reactor
model is shown in Figure 1. Although the reactor model
is integrated in a simulation of the whole plant, this work
focuses explicitly on the reactor model and other
components are not considered. The presented reactor
yields are therefore ideal yields supposing perfect product
separation.

The char fraction is modelled by a series of fractional


char species, which allows including H, O, N, S and Cl in
the char fraction, accounting for the real pyrolysis char
composition which contains an important share of oxygen
(influencing the heating value of the char). Nevertheless,
as these fractional species pose problems in subsequent
Aspen Plus blocks due to missing properties and/or
wrongly estimated properties, these are all converted in
non-conventional char in the following secondary
pyrolysis reactor (RCTSECND in Figure 1).

Figure 1: Aspen Plus flowsheet of the reactor model


3.1. Decomposition reactor
The decomposition reactor (RDECOMP in Figure 1)
is responsible for the decomposition of the biomass
(modelled as a non-conventional component) into its
fractional components: cellulose, hemicellulose, and
lignin. Hemicellulose and cellulose are represented in the
simulation by their monomers: C5H8O4 (xylan) and
C6H10O5 (xylose-like cellulose monomer), respectively.
Lignin is represented by 6 different monomers with
different O/C and H/C ratios, which permits adjusting the
atomic composition of the decomposition products to the
biomass composition by varying the amounts of the
different lignin components [9]. The nitrogen content of
the biomass is taken into account by including two
representative N-containing species in the decomposition
products, glutamic acid and pyrrole, again with different
O/C and H/C ratios to adapt to different biomass
compositions. Both are frequent in biomass [10], the
amino acid represents proteins while pyrrole is the basic
compound of more complex molecules like chlorophyll
or porphyrins.
The decomposition reactor is simulated using an
RYield reactor, for which a calculation block provides
the yield data for the decomposition fractions. The
fractions are calculated by an iterative algorithm that, for
a given initial atomic and fractional composition of the
biomass, adjusts the lignin components to fit the atomic
composition.
3.2. Primary pyrolysis reactor
The primary pyrolysis reactor (RCTPRIM in Figure
1) is alternatively an RBatch or RCStir reactor which
calculates the pyrolysis products based on a kinetic
reaction mechanism. The reaction mechanism is based
primarily on the works of Faravelli et al. [9], Dupont et
al. [11], Calonaci et al. [12], Ranzi et al. [13], van de
Velden et al. [14], and Graham et al. [15]. It implements
over 150 individual reactions, including primary
decomposition, secondary decomposition and radical
substitution.
The reactor type can be chosen according to the
pyrolysis reactor which wants to be modelled. For fast
pyrolysis, an RCStir reactor is used, while an RBatch
reactor is used for slow pyrolysis. For modelling different
reactor types, the operating temperature and the bed and
vapour residence times for the simulated reactor are
required.

3.3. Secondary pyrolysis reactor


Apart from the homogeneous vapour-phase reactions,
heterogeneous tar condensation reactions occur in the hot
vapour phase. These are partially catalysed by the alkali
components in the char fraction [5,16,17] and give rise to
increasing char and gas yields at the expense of the
organic liquid fraction. Especially for biomass with high
alkali ash content and at longer residence times play these
effects an important role in the overall reaction
mechanism. As these condensation reactions are very
complex and not yet fully understood, no chemical
models for their detailed description can be found in
literature. To account for them nevertheless, a linear
regression algorithm based on works which give
quantitative data about the change in fractional yields
[16] has been included in the reactor model. For this
purpose, an RYield reactor (RCTSECND in Figure 1) is
implemented that modifies the fractional yields according
to a linear regression depending on vapour residence
time, temperature and char alkali metal content.
Furthermore, this reactor condenses all the different char
fractions into one non-conventional component (NCChar) with the corresponding elemental composition. The
char shows an atomic composition similar to the values
found in literature. Radicals and other non-ordinary
fractional components still contained in the oil are
condensed into conventional model components. This
significantly simplifies the simulation as the resulting
bio-oil then consists only of standard components with
well-defined properties.
3.4. Product gas recycle loop
For modelling bubbling fluidised bed reactors and
other types that require a fluidising gas, an optional
product gas recycle loop (RFLUEGAS in Figure 1) is
implemented. This is important as a high amount of
fluidising gas changes the concentrations of the
components in the reactors, which might affect the
results.
3.5. Hot sand loop
The pyrolysis reactor requires a certain amount of
heat. This heat is provided by a combustor which burns
the pyrolysis gases and part of the char (combustion
reactor not included in Figure 1). Hot sand is circulated
between the combustor and the pyrolysis reactor
(SANDCYCL in Figure 1) and acts as a heat carrier. It is
heated (HXCCOMB) and then recycled to the pyrolysis
reactor where it heats the biomass particles by direct
contact in the reactor bed (HXCREACT). For simplicity
purposes, the sand loop is separated in the simulation and
the heat is transferred by a heat exchanger to the reactor
bed, although, especially in circulating fluidised bed
reactors, it is part of the reactor product flow and not
separated until the cyclone.

4. SIMULATION RESULTS
In order to validate the reactor model, numerous
simulation runs were performed and the results were
compared with literature data.
4.1. Temperature and time dependency
The dependency of the fractional yields on
temperature and residence time is tested with two
different types of feedstock: pine wood and wheat straw.
The two types of feedstock are chosen due to their
different properties, with wheat straw being a rather lowlignin biomass with a high content of ash and trace
elements, while pine wood is a rather pure biomass
with very low ash and trace element content. The results
of the simulation runs are presented in Figures 2-7
(temperature and time axes are not linear).

The effect of the feedstock composition can be


observed when comparing the yield curves for the two
feedstock types. Higher liquid yields of above 75% are
achieved by the pine wood feedstock, while wheat straw
hardly exceeds 65% liquid yield. This is mainly due to
the higher ash content of the straw, which catalyses
secondary decomposition reactions and increases gas and
char yields at the expense of the liquid fraction. Char
gasification reactions are not included in the model, so
char yields tend to be somewhat high for high
temperatures and long residence times, and the results for
these operation conditions should be taken with care.
Nevertheless, these are rather extreme conditions which
do not represent the typical operation range of pyrolysis
reactors for lignocellulosic biomass.

Figure 5: Gas yield with wheat straw as feedstock


Figure 2: Gas yield with pine wood as feedstock

Figure 6: Oil yield with wheat straw as feedstock


Figure 3: Oil yield with pine wood as feedstock

Figure 7: Char yield with wheat straw as feedstock


Figure 4: Char yield with pine wood as feedstock
As can be observed, the yield curves show the typical
shape for pyrolysis reactions, with a maximum liquid
yield at around 500C and residence times of 1 s [1820].
At low temperatures and short residence times, the char
yield increases sharply, indicating incomplete
decomposition.

4.2. Product composition


In order to assess the impact of the residence time not
only on the fractional yields, but on the pyrolysis product
composition, four different types of feedstock are
simulated under fast (500C, 1 s residence time) and slow
pyrolysis conditions (425C, 30 min residence time,
heating rate 0.5C/s). The results of the simulation runs
are shown in Figures 8 and 9.

The influence of the catalytic effect of the alkaline


metal contained in the ashes can be observed for wheat
straw, whose char yield is higher than that of lower ashcontaining woody feedstocks, especially under slow
pyrolysis conditions.
Furthermore, the effect of the feedstock type on the
bio-oil fractions can be seen. The higher cellulose and
hemicellulose content of, for example, wheat straw in
comparison to pine wood gives rise to a lower share of
degraded lignin compounds and more organic acids
contained in the bio-oil.
In the slow pyrolysis oils the water content is
significantly higher than in the fast pyrolysis oils, as the
secondary condensation reactions increase water and char
yields. A decrease in the degraded lignin and the sugarderived fraction can also be observed, along with an
increase in the water fraction, indicating the
decomposition / polymerisation of these fractions which
is more severe in the ash-rich wheat straw oil [16,21].

As can be observed, the correlation between the product


compositions given by the simulation and the literature
data is high. The proposed simulation gives somewhat
lower water content, but on the other hand it accounts for
organic acids which are not explicitly analysed in the
used literature data.

Figure 8: Product yields for fast pyrolysis of four


different types of feedstock

Figure 11: Composition of pine wood fast pyrolysis oil Literature data [23]

Figure 9: Product yields for slow pyrolysis of four


different types of feedstock
4.3. Verification with literature data
Verification with literature data is difficult, as works
providing all the required data (details about feedstock
composition
regarding
fractional
and
atomic
composition, and a detailed analysis of the fractional
composition of the bio-oil) are hard to find. Nevertheless,
verification with works that do not give all required
information [2225] is possible using standard values
regarding biomass composition from the Phyllis database
[26]. Figures 10-13 show the comparison of the
simulation-derived composition of the bio-oils from fast
pyrolysis of two types of feedstock with the composition
given in literature [23]. Pine wood and eucalyptus are
used as only for these all required information is given.

Figure 10: Composition of pine wood fast pyrolysis oil Simulation results

Figure 12: Composition of eucalyptus wood fast


pyrolysis oil - Simulation results

Figure 13: Composition of eucalyptus wood fast


pyrolysis oil - Literature data [23]

CONCLUSIONS

Overall, the results obtained from the simulation for


the different (woody and herbaceous) feedstocks are
found to be in line with existing studies in the
temperature range typical for pyrolysis. The model gives
temperature-yield curves comparable to the ones
published in literature and the dependency of the fraction
yields on residence time and feedstock composition
correspond. Data for rigorous verification providing all
required input data and at the same time a detailed
product analysis is scarce but could be done for two types
of feedstock under fast pyrolysis conditions. Results also
show good correlation between experimental and
simulation data under this regard. The simulation gives a
bio-oil of slightly lower water content, but on the other
hand accounts for organic acids which are not considered
in the available literature data. In conclusion, the model is
found to be suitable for predicting fast and slow pyrolysis
reactions for lignocellulosic biomass feedstock and can
be used for calculating pyrolysis products without the
need for experimental data.
6

REFERENCES

[1] EC. Proposal for a Directive of the European


Parliament and of the Council amending Directive
98/70/EC relating to the quality of petrol and diesel
fuels and amending Directive 2009/28/EC on the
promotion of the use of energy from renewable
sources. Brussels: European Commission (2012).
[2] EEA. How much bioenergy can Europe produce
without harming the environment? Copenhagen
(2006).
[3] Bridgwater A. Fast pyrolysis processes for biomass.
Renewable and Sustainable Energy Reviews
(2000);4:173.
[4] Bridgwater AV. Biomass Fast Pyrolysis. Thermal
Science (2004);8:2149.
[5] Venderbosch R, Prins W. Fast pyrolysis technology
development. Biofuels, Bioproducts and Biorefining
(2010);4:178208.
[6] Bridgwater AV. Upgrading biomass fast pyrolysis
liquids. Environmental Progress & Sustainable
Energy (2012);31:2618.
[7] Xiu S, Shahbazi A. Bio-oil production and
upgrading research: A review. Renewable and
Sustainable Energy Reviews (2012);16:440614.
[8] Meier D, van de Beld B, Bridgwater AV, Elliott DC,
Oasmaa A, Preto F. State-of-the-art of fast pyrolysis
in IEA bioenergy member countries. Renewable and
Sustainable Energy Reviews (2013);20:61941.
[9] Faravelli T, Frassoldati A, Migliavacca G, Ranzi E.
Detailed kinetic modeling of the thermal degradation
of lignins. Biomass and Bioenergy (2010);34:290
301.
[10] Hansson K-M, Samuelsson J, Tullin C, mand L-E.
Formation of HNCO, HCN, and NH3 from the
pyrolysis of bark and nitrogen-containing model
compounds. Combustion and Flame
(2004);137:26577.
[11] Dupont C, Chen L, Cances J, Commandre J-M,
Cuoci A, Pierucci S, et al. Biomass pyrolysis:
Kinetic modelling and experimental validation under
high temperature and flash heating rate conditions.
Journal of Analytical and Applied Pyrolysis
(2009);85:2607.

[12] Calonaci M, Grana R, Barker Hemings E, Bozzano


G, Dente M, Ranzi E. Comprehensive Kinetic
Modeling Study of Bio-oil Formation from Fast
Pyrolysis of Biomass. Energy & Fuels
(2010);24:572734.
[13] Ranzi E, Cuoci A, Faravelli T, Frassoldati A,
Migliavacca G, Pierucci S, et al. Chemical Kinetics
of Biomass Pyrolysis. Energy & Fuels
(2008);22:4292300.
[14] Van de Velden M, Baeyens J, Brems A, Janssens B,
Dewil R. Fundamentals, kinetics and endothermicity
of the biomass pyrolysis reaction. Renewable Energy
(2010);35:23242.
[15] Graham RG, Bergougnou MA, Freel BA. The
kinetics of vapour-phase cellulose fast pyrolysis
reactions. Biomass and Bioenergy (1994);7:3347.
[16] Hoekstra E, Westerhof RJM, Brilman W, Van Swaaij
WPM, Kersten SRA, Hogendoorn KJA, et al.
Heterogeneous and homogeneous reactions of
pyrolysis vapors from pine wood. AIChE Journal
(2012);58:283042.
[17] Park H, Park Y, Dong J, Kim J, Jeon J, Kim S, et al.
Pyrolysis characteristics of Oriental white oak:
Kinetic study and fast pyrolysis in a fluidized bed
with an improved reaction system. Fuel Processing
Technology (2009);90:18695.
[18] Di Blasi C, Signorelli G, Di Russo C, Rea G. Product
Distribution from Pyrolysis of Wood and Agricultural
Residues. Industrial & Engineering Chemistry
Research (1999);38:221624.
[19] Neves D, Thunman H, Matos A, Tarelho L, GmezBarea A. Characterization and prediction of biomass
pyrolysis products. Progress in Energy and
Combustion Science (2011);37:61130.
[20] Bridgwater AV. Review of fast pyrolysis of biomass
and product upgrading. Biomass and Bioenergy
(2012);38:6894.
[21] Kawamoto H, Murayama M, Saka S. Pyrolysis
behavior of levoglucosan as an intermediate in
cellulose pyrolysis: polymerization into
polysaccharide as a key reaction to carbonized
product formation. Journal of Wood Science
(2003);49:46973.
[22] Horne PA., Williams PT. Influence of temperature on
the products from the flash pyrolysis of biomass. Fuel
(1996);75:10519.
[23] Oasmaa A, Solantausta Y, Arpiainen V, Kuoppala E,
Sipil K. Fast Pyrolysis Bio-Oils from Wood and
Agricultural Residues. Energy & Fuels
(2010);24:13808.
[24] Faix A, Schweinle J, Schll S, Becker G, Meier D.
(GTI-tcbiomass) life-cycle assessment of the BTOprocess (biomass-to-oil) with combined heat and
power generation. Environmental Progress &
Sustainable Energy (2010);29:193202.
[25] Bajus M. Pyrolysis of woody material. Petroleum &
Coal (2010);52:20714.
[26] ECN-Biomass. Phyllis Database n.d.
http://www.ecn.nl/phyllis/. Last visited 2013-04-25
7

ACKNOWLEDGEMENTS

This research has been partly supported by the Regional


Government of Madrid (S2009/ENE-1743) and the
Spanish Ministry of Economy and Competitiveness
(ENE2011-29643-C02-01 and IPT-2012-0219-120000).

Potrebbero piacerti anche