Sei sulla pagina 1di 18

Chemical Engineering Science 148 (2016) 14–31

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Review

Review of gasification fundamentals and new findings: Reactors,


feedstock, and kinetic studies
Nader Mahinpey n, Arturo Gomez
Department of Chemical and Petroleum Engineering, Schulich School of Engineering, University of Calgary, Calgary, Canada AB T2N 1N4

H I G H L I G H T S

 The presence of maximum gasification rate is not related to char surface morphology.
 Activation energy estimated at different conversions suggests reaction mechanism.
 Separation of pyrolysis and gasification experiments affects the kinetics.
 Most important gasification variables are char surface area and alkali content.
 The random pore model (RPM) does not represent the real gasification mechanism.

art ic l e i nf o a b s t r a c t

Article history: The purpose of this work is to comprehensively review recent findings on gasification kinetics related to
Received 19 October 2015 the reaction mechanism, the effects of the experimental procedure on the overall reaction, and methods
Received in revised form to evaluate kinetic models.
22 March 2016
Understanding the recent findings in gasification requires the presentation of selected publications
Accepted 24 March 2016
Available online 28 March 2016
that cover several potentially ambiguous concepts. Firstly, conventional reactor definitions are not suf-
ficient given that three reactions (combustion, pyrolysis and gasification) occur in an industrial gasifier.
Keywords: Secondly, variables affecting char reactivity during gasification show that the char surface area is the
Gasification rate most complex variable to analyze. Also, single-step kinetic models provide better descriptions of the
Kinetic model reaction mechanism assuming kinetic control of the overall reaction.
Random pore model
The new findings are related to the gasification reaction mechanism and the interpretation of kinetic
Mass transfer
results: (a) the maximum gasification rate observed is a consequence of the gas concentration devel-
Char surface development
Activation energy
opment upon changing the inert gas for the gasifying agent, and not due to changes in the char surface
during gasification, as proposed by Bhatia and Perlmutter in 1980; (b) increasing the time of holding the
char at the reaction temperature in an inert gas decreases the gasification rate; (c) char micropore surface
area and alkali content are the two most important variables affecting the gasification rate; and,
(d) laboratory studies may not represent the real gasification mechanism under kinetically controlled
conditions due to the presence of mass transfer limitations from interparticle diffusion. Therefore, the
acceptance of intrinsic kinetics reported in the literature should be revalidated.
& 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2. Gasification process and reactors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1. Gasification applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2. Gasifying reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3. Gasifier types and stream configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Abbreviations: EFR, entrained flow reactor; FBR, fluidized-bed reactor; ICM, integrated core model or power-law model; LH, Langmuir–Hinshelwood; NDF, normal dis-
tribution function model; SCM, shrinking core model; TGA, thermogravimetric analyzer; RPM, random pore model; VM, volumetric model
n
Corresponding author.
E-mail address: nader.mahinpey@ucalgary.ca (N. Mahinpey).

http://dx.doi.org/10.1016/j.ces.2016.03.037
0009-2509/& 2016 Elsevier Ltd. All rights reserved.
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 15

2.3.1. Downdraft and updraft configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18


2.3.2. Feedstock distribution into the reactor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3.3. Solid by-product disposal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3. Variables affecting char reactivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.1. Reaction temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2. Total pressure and gasifying agent partial pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3. Pyrolysis and char surface area development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4. Mineral content and composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4. Kinetic studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1. CO2 gasification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2. Steam gasification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3. Assumptions and simplifications within the literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3.1. Mass transfer limitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3.2. Char synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3.3. Existence of a maximum reaction rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.4. Chemical reaction kinetic models for char gasification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.5. Langmuir–Hinshelwood kinetic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5. New findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.1. Real nature of the maximum gasification rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.2. Effect of the isothermal pyrolysis in char reactivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.3. Most significant variables affecting char reactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.4. Estimation of kinetic parameters independent of the kinetic model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.5. Reduction of interparticle diffusion in experimental studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1. Introduction IGCC technology involves two major processes: coal gasification


and combined cycle energy generation (Hoffmann and Szklo,
The instability of oil prices has motivated research into the use 2011); however, its feasibility is significantly affected by the con-
of coal and other solid carbonaceous feedstocks as energy sources. sumption of natural gas.
Current coal combustion technology was developed for high-grade Coal and natural gas are the two major fossil fuels used in
coals (related to high heat content per kg of coal) and is still the electricity generation; therefore, an increase in the consumption of
dominant process for thermal power generation. However, its use one means a decrease in the consumption of the other, when there
has been restricted worldwide due to its higher impact on an- is a surplus of these fuels. In the spring of 2012, natural gas prices
thropogenic emissions causing global climate change (Matthews dropped to historically low levels, which led to a significant in-
et al., 2009) and environmental concerns, such as high particulate crease in natural gas-fired generators being used over coal-fired
matter and sulfur dioxide (SO2) pollution. On the other hand, high- generators. In early 2013, coal regained market share after natural
grade coals are not abundant and technologies to obtain ash-free gas prices rose, which led natural-gas-fired plants to losing their
coal, such as solvent extraction, are necessary (Rivolta Hernandez competitive advantage (EIA, 2013). Natural gas consumption also
et al., 2012). In spite of these concerns, coal accounts for a major has two peaks each year: the first one is during the winter, when
share of the electrical power generated in the western and central cold weather increases the demand for natural gas heating; and,
provinces of Canada (Thibault, 2012). the second peak is during the summer, when hot weather creates
The terminology used for coal has been extended to encompass demand for air conditioning (EIA, 2011). Significant uncertainty
all solid carbonaceous feedstock, including non-conventional fuels, remains about the future prices of coal and natural gas. None-
such as petroleum coke (petcoke). There has been an increase in theless, higher prices of coal give natural-gas-fired plants a com-
petcoke production in Alberta, which has been generated as a by- petitive advantage and vice versa (EIA, 2013).
product of oil sands upgrading. Petcoke may be used as feedstock Gasification is an alternative for generating a clean fuel for
to replace natural gas in oil sands extraction and upgrading heating purposes. The first industrial application was the pro-
(Cousins et al., 2012). Other solid fuels with high ash contents and duction of syngas for lighting in 1792; by the late 1920s, there
similar or lower heat capacities are low-grade fuels, such as bio- were more than 1200 operational gasification plants in the United
mass. High-quality coals are commonly associated with high car- States (Rezaiyan and Cheremisinoff, 2005). Plasma gasification is
bon and low mineral contents (O'Keefe et al., 2013). Low-rank useful in the treatment of municipal solid waste, and it can be
coals exhibit a broad spectrum of properties directly associated integrated with power generation using a combined cycle (IPGCC)
with less maturity and higher inorganic matter (Bielowicz, 2012; (Materazzi et al., 2014; Ruj and Ghosh, 2014). It also serves as an
Cousins et al., 2012; O'Keefe et al., 2013). The coal rank is de- alternative for producing syngas, which is the feedstock for syn-
termined by carbon, moisture, volatile matter content, and ca- thetic fuel manufacturing and hydrogen production for refinery
lorific value; and, the general rule is that high-rank coals are and fuel upgrading (Cousins et al., 2012; Dermot, 2013).
harder materials with less moisture, less volatile matter, and Researchers around the world have been investigating the ki-
higher calorific values (Schweinfurth, 2009). netics of gasification in order to increase the gasification rate at
The utilization of low-grade feedstocks is possible using gasi- lower temperatures, which is critical to its implementation.
fication instead of combustion; however, industrial-scale im- However, experimental results have not shown uniform trends,
plementation poses challenges (NETL, 2002). The most successful and it is difficult to compare kinetic parameters of the same
demonstration of gasification is the use of the integrated gasifi- feedstock produced from different experimental setups. The di-
cation combined cycle (IGCC) in power generation. The established verse information presented in the literature with inconsistent
16 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

Nomenclature R ideal gas law constant (kJ mol  1 K  1)


S specific surface area (m2 g  1)
EA activation energy (kJ mol  1) t time (min)
f(X) function of the conversion in a general rate law T temperature (K)
(dimensionless) X conversion (dimensionless)
G(X) integrated form of dX/f(X) (dimensionless)
ΔHR, i heat of reaction ‘i’ at reaction conditions (kJ mol  1) Greek letters
k rate constant (min  1)
ko frequency factor (min  1) α intercept of the logarithm of time (min) versus re-
kM rate constant for the particular kinetic model ‘M’ ciprocal of temperature (K  1)
(min  1) π mathematical constant, approx. 3.14159
ma mass of ash (g) Δ gradient of a determined variable
mo initial mass of char (g) ψ parameter of the random pore model describing char
mt mass of char at the particular time ‘t’ (g) surface (dimensionless)
n reaction order for the integrate core model
r reaction rate (min  1)

criteria makes it difficult to extrapolate results or correlate in- 2012a). The reduction of water (H2O) in steam gasification is the
dependent studies. In contrast, the scientific community has been most effective way of increasing H2 production (Kim et al.,
more focused on theoretical research, particularly chemical reac- 2014). CO2 promotes the Boudouard reaction (endothermic) to
tion models with non-unique trends on the estimation of kinetic produce CO (Hernandez et al., 2012b). CO2 may also be recircu-
parameters (Di Blasi, 2009). These gaps must be addressed to lated with O2 within oxy-fuel combustion/gasification (Gil et al.,
consistently use the results of laboratory-scale kinetic studies for 2012).
the design of industrial gasifiers.
The goal of this review is the clarification of gasification topics Some limitations of incineration include harmful emissions of
that have been ambiguously mentioned in the literature. This re- acidic gases and volatile organic compounds, as well as the pro-
view includes recent findings about gasification experimental duction of toxic wastes, such as dioxins, which contribute to air
studies and kinetic modeling, which may help researchers to pollution (Ma et al. 2011). Compared with incineration and com-
comprehensively understand gasification with a minimum bustion of the solid fuel, the use of syngas produces less solid
amount of assumptions. New concepts clarify why the separation waste and emits fewer harmful gases into the environment
between gasification and pyrolysis can induce misinterpretations (Duchesne et al., 2012). Syngas combustion is an efficient and
about the reaction mechanism. clean technology due to low emissions of mono-nitrogen oxides
(NOx) and sulfur oxides (SOx) and has a high rate of energy re-
covery (Hernandez et al., 2012c; Ma et al., 2011). Gasification, on
2. Gasification process and reactors the other hand, allows the installation of small, low-cost, efficient
reactors to reduce storage and transport costs. It allows the re-
Gasification is the thermochemical conversion of a carbon- covery of available energy from low-value materials, such as bio-
based feedstock into a combustible product gas with the use of a mass wastes and low-rank coals, reducing disposal costs and en-
gasifying agent or agents, such as steam and carbon dioxide (CO2) vironmental concerns (Hernandez et al., 2012c). Gasification also
(Ma et al., 2011). The main product gases – carbon monoxide (CO) provides an alternative for solid animal waste disposal through
and hydrogen (H2) – are used to produce heat and electricity or are manure-fuel gasification (Pian, 2002).
converted into liquid hydrocarbons (Bremaud et al., 2005; Coda
et al., 2007). The non-reactive components, called ash, are not 2.1. Gasification applications
totally inert, because they can change their properties due to solid-
phase changes and partial oxidation or reduction. There are various advantages to gasifying low-grade fuels over
Industrial gasification consists of three important processes conventional processes and fossil fuels. Gasification processes can
(Fermoso et al., 2008; Qian et al., 2013; Zhao et al., 2009): be applied to convert wastes to valuable forms of energy. These
wastes range from any type of biomass and forestry residues to
 Pyrolysis reactions: This set of complex reactions involves petcoke. This ability to convert waste is beneficial for the en-
heating solid fuel, usually up to 700 °C, to produce char and vironment by preserving lands and providing an economic ad-
release volatiles. Tars may also be produced when volatiles li- vantage in terms of saving storage costs, because gasification uses
quefy at low temperatures. straw, agricultural waste, and other products that cannot be eco-
 Oxidation reactions: Char and volatiles are combusted with nomically processed using other technologies.
oxygen (O2) to produce the necessary gasifying agents (steam Gasification can be combined with power generation using IGCC
and CO2) and CO. This reaction is exothermic, and the released power plants to efficiently and cleanly generate energy. Large
heat is used for reduction reactions. companies, such as Siemens, GE, ConocoPhillips, and Shell, have
 Reduction reactions (mainly gasification reactions): Char, tar, built IGCC facilities around the globe (Hoffmann and Szklo, 2011).
and hydrocarbons are gasified with CO2 and steam to produce IGCC plants are designed with sulfur and CO2 capture systems to
synthetic gas (syngas), which is mainly composed of CO, H2, and remove most of these harmful constituents from the produced gas.
CH4. These reactions are endothermic and require the heat As shown in Fig. 1, one of the possible gasification routes is power
produced from the prior oxidation reactions. Steam promotes generation using different feedstocks; however, chemical synthesis
the steam reforming (endothermic) of char and tar, as well as is also achievable with syngas through the use of the Fischer–
water–gas shift reactions (exothermic) (Hernandez et al., Tropsch (FT) process (Bremaud et al., 2005; Coda et al., 2007).
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 17

Coal
Lignite Tar
Feedstock removal IGCC
Petcoke
treatment
Biomass
Municipal waste FT
CO shift
Syngas Synthesis
Gasifier
Sulfur Methanol
Recycled CO2 Gasifying removal Synthesis
Superheated Steam agent
and
Air reaction CO2 Ammonia
O2 medium removal Production

Ash

Fig. 1. Possible plant configurations for different gasification applications.

A general view of this technology implies an understanding of et al., 2014). CO2 promotes the Boudouard reaction (endothermic)
the following: to produce CO (Hernandez et al., 2012b). CO2 may also be com-
bined with O2 for oxy-fuel combustion prior to the gasification
 Raw materials: gasification is a technology that can be im- stage, which is a more efficient process than using air as an oxi-
plemented with liquid feedstock; however, this provides no dant due to reduced nitrogen (N2) dilution (Gil et al., 2012).
economic advantage. Solid fuels are the cheapest fuels, but their The most important reactions considered for kinetic studies,
transport is expensive; therefore, they can be easily handled if excluding combustion, are presented in Eqs. (1)–(5) at standard
they are converted into a gas stream. conditions, indicating which reactions are reversible in the tem-
 Gasification process: there are two important elements that perature range for which gasification is kinetically controlled. The
constitute the baseline for all possible configurations: Boudouard reaction is the controlling step at low gasification
temperatures during CO2 gasification (lower than 1000 °C)
– Reactor type: the reactor influences the way that the solid is (Ahmed and Gupta, 2011; Tremel and Spliethoff, 2013).
held in the reaction chamber and the direction of the gas flow
C þCO2⇆2CO ΔHR1 ¼ 172.5 kJ/mol (1)
with respect to the feed of the solid. There are multiple chemical
reactions taking place in an industrial gasifier, where the se-
quence of reactions is determined by the reactor configuration. C þH2O-H2 þCO ΔHR2 ¼131.3 kJ/mol (2)

 Heat transfer and heating sources: the main mechanism is ra- COþH2O⇆H2 þ CO2 ΔHR3 ¼  41.2 kJ/mol (3)
diation, because gasification is a set of endothermic gas–solid
reactions occurring at temperatures above 500 °C. For this rea- C þ2 H2-CH4 ΔHR4 ¼  74.5 kJ/mol (4)
son, indirect heating is not efficient when the size of the reactor
increases. The most common gasifiers are directly heated by CH4 þH2O⇆COþ 3 H2 ΔHR5 ¼205.8 kJ/mol (5)
overheated steam and partial combustion of the feedstock.
Above 700 °C, the Boudouard reaction takes place, i.e., gasifi-
 Syngas quality: tar and other compounds must be removed to
cation in Eq. (1). If the heating rate is higher than 200 °C/min,
use the syngas (mainly CO and H2). Without syngas cleaning,
there is no weight loss difference during the pyrolysis step, re-
there are no environmental advantages to gasification with re-
gardless of the sweeping gas used (Silbermann et al., 2013). The
spect to direct combustion. The main advantage of gasification
concept of the introduction of an isothermal step to separate
is the ability to use smaller equipment since this is a pre-
pyrolysis from gasification is not the best solution, as pyrolysis is
combustion treatment.
much faster than gasification and significantly affects char re-
 Final application: IGCC technology is the best practical example of
activity (Tremel and Spliethoff, 2013; Zhu et al., 2008).
gasification implementation. There are other options, such as che-
CO2 gasification and steam gasification (Eqs. (1) and (2), re-
mical synthesis, where the most important factor is the CO/H2 ratio.
spectively) are endothermic reactions; when they are combined,
they improve the H2 yield at higher temperatures. The water–gas
2.2. Gasifying reactions shift reaction (Eq. (3)) increases the H2/CO ratio in the final mix-
ture, but does not greatly affect the heating value of the syngas
The gasifying agent is the gas that is required to perform ga- because H2 and CO combustion heats are almost identical on a
sification reactions. This gas reacts with chars, tars, and gases to molar basis. The variation of the heat of reaction under various
produce syngas. Some researchers have referred to air, O2, steam, gasification temperatures is smaller than 5 kJ/mol for all hetero-
and CO2 as gasifying agents. However, only steam and CO2 are geneous reactions; on the other hand, both the homogeneous re-
used in gasification reactions; oxygen (O2) is applied in the com- actions of, water–gas shift (Eq. (3)) and steam methane reforming
bustion step to provide the required energy for the endothermic (Eq. (5)), change from  41 kJ/mol and 206 kJ/mol at standard
reactions. As the heat source is an important attribute at industrial conditions to  33 kJ/mol and 227 kJ/mol at 900 °C, respectively.
scale, the configuration of reactor becomes an essential factor Combustion is important; however, under low air intake or pure
when O2 is used as oxidizing agent. O2 for oxy-fuel systems, the gasifier operates under reduction con-
Steam and CO2 have their own advantages in gasification. ditions. CO2 is usually produced in situ by partial combustion of the
Steam promotes the steam reforming (endothermic reaction) of char, but its gasification efficiency is low when using an entrained
char and tar, as well as the water–gas shift reaction (exothermic) flow reactor (EFR), because residence time in the order of a fraction
(Hernandez et al., 2012a). The reduction of H2O in steam gasifi- of a second is not sufficient to ensure that the Boudouard reaction
cation is the most effective way of increasing H2 production (Kim takes place. Industrial gasifiers can be ideally considered as three
18 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

reactors connected in series for combustion, pyrolysis, and gasifica- reactions is part of the model and when kinetics is considered in-
tion; however, the complexity increases when the sequence of these stead of thermodynamic equilibrium.

2.3. Gasifier types and stream configurations

The sequence of reactions, gasifying agents, and stream con-


figurations determine the type of reactor. The main difference
between reactor types is the sequence of reactions for solid and
gas species within the gasifier. For some gasifier types, the steps
for solid feed are drying, pyrolysis, char gasification, and ash
melting. Fig. 2 shows a schematic of the plasma gasification pro-
cess, which has the same elements as conventional gasification
with the addition of continuous melting of the ash (Zhang et al.,
2013).

2.3.1. Downdraft and updraft configuration


When flow configuration is considered, co-current and counter-
current reactors are the equivalents of downdraft and updraft gasi-
fiers, respectively. The sequence of reactions presented in Fig. 2 is one
of many possible configurations, given the position of the feedstock
and gasifying agent inlets. Fig. 3 presents only the main set of gas–
solid chemical reactions that occur in a gasifier according to the
stream configuration: combustion, pyrolysis, and gasification.
Fig. 3 illustrates that the order of the chemical reactions de-
pends on the position of the gas or solid inlet. The position has
important consequences:

 Tar formation is reduced using a downdraft reactor, because


combustion follows pyrolysis (Ruiz et al., 2013). This result is
impossible to achieve in an updraft reactor, and tar removal is
required prior to the use of gas products.
 Residence time is shorter in the downdraft reactor, because
solids move faster due to gravity and drag force being aligned in
the same direction. Lower carbon conversion is achieved in this
configuration compared with a similar size updraft reactor. This
may be a problem if the solid product is not treated or if the
reactor size is not increased. If the reactor is packed, its con-
version can be the same or higher as that of the updraft reactor;
however, the heating value of the gas products is reduced
compared to other configurations (Kihedu et al., 2014).

Fig. 2. Schematic of plasma gasification process (Reprinted with permission from


2.3.2. Feedstock distribution into the reactor
Zhang et al. (2013)).
Fixed and fluidized bed reactors are the most common reactor

(a) Downdraft Gasifier (b) Updraft Gasifier

Feedstock
Feedstock Gasifying agent + Gas products
+ catalyst ?
catalyst? (1) Pyrolysis (3) Pyrolysis

(2) Combustion

(2) Gasification

(3) Gasification

(1) Combustion
Gas products
Gasifying agent
Ash / slag Ash / slag
Fig. 3. Configuration of gasifiers according to direction of the feed streams: (a) downdraft, (b) updraft. Three sets of chemical reactions occur in a gasifier: (1) combustion,
(2) pyrolysis, and (3) gasification.
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 19

Counter-currrent Co-ccurrent
(Downdraaft) (U p
pdraft)

Fig. 5. Fluidized bed reactors used for gasification. (Reprinted with permission
from Warnecke (2000)).

Fig. 4. Fixed bed reactor configurations. (Reprinted with permission from War- and bed material are returned to the reaction vessel. In contrast to
necke (2000)).
the bubbling design, the circulating fluidized bed has a very high
gas velocity and can recycle very large amounts of solids; there-
types, but other types, such as the EFR, have been widely used for fore, it is capable of handling large feed throughputs (McKendry,
gasification. 2002). Bubbles do not exist in the circulating design; hence, oxy-
gen cannot be trapped and efficiency does not decrease (Ruiz et al.,
2.3.2.1. Fixed bed reactor. Fixed bed reactors are sometimes called 2013).
moving beds, due to the slow motion of the solid with respect to The operating temperature of FBR gasifiers ranges from 800 to
the gas (Ruiz et al., 2013). Fig. 4 illustrates the two possible con- 1000 °C, which prevents the buildup of ash. Therefore, these re-
figurations for this reactor type: updraft (co-current) and down- actors have no limitations in processing feed with high ash content
draft (counter-current) reactors (Warnecke, 2000). The particular in this temperature range (Ruiz et al., 2013). The main operational
type of reactor is determined by the relative position of the intake difficulty is the potential of slagging at low temperatures when
for the solid feedstock and gasifying agent. using biomass, which contains ash that melts at temperatures
In the downdraft configuration, there is a lower content of tar below 1000 °C (McKendry, 2002; Warnecke, 2000). The main
(Ruiz et al., 2013) and a reduction in the heating value of the consideration is the removal of ash from this type of reactor, which
product gas (Kihedu et al., 2014) than with the updraft reactor. In has mostly been ignored in the reported gasification studies with
addition, the downdraft configuration is less flexible than updraft bubbling FBRs but its effects have been reported in recent studies
reactors, due to the lack of operational versatility of different about agglomeration and defluidization in FBRs (Datta et al., 2015;
feedstocks (Ruiz et al., 2013). Lu et al.; 2015; Serrano et al., 2015).
The most significant advantage of fixed bed reactors is their
high conversion rate for a given volume of reactor; however, they 2.3.2.3. Entrained flow reactor (EFR). Different authors have con-
are usually limited to operations at temperatures below 1000 °C sidered the use of EFRs in coal gasification (Collot, 2006; Tremel
(Warnecke, 2000). et al., 2013). Gasification with an EFR has advantages such as
flexibility, low tar content of the gas product, and high char con-
2.3.2.2. Fluidized bed reactor. Fluidized bed reactors (FBRs) have version. These benefits are the reasons they are the most widely
been extensively applied in coal gasification for several years used type of gasifier; over 75% of the 160 gasification projects in
(McKendry, 2002). However, their application is limited, due to 2005 used EFR (Minchener, 2005). In EFR gasifiers, usually both
low carbon conversion (Tremel et al., 2013). This reactor type is the feed powders (less than 75 μm) and gasifying agents enter
continuously fed close to the bottom, and the gasifying agent flows from the top of the gasifier. The speed of the gas through the re-
in continuously from the bottom. The main advantage of FBRs is actor is fast enough to entrain all the feed particles.
better distribution, thereby reducing heat and mass transfer lim- As the EFR requires a very small particle size, grinding and
itations. However, this reactor type is equivalent to a combination achieving the appropriate feed size is critical and can sometimes
of downdraft and updraft reactors, and the associated problem of be challenging. This reactor normally operates at temperatures
low tar cracking must be considered when using an FBR. between 1200 °C and 1600 °C and at pressures between 2 and
As shown in Fig. 5, there are two common types of FBRs: the 8 MPa (McKendry, 2002; Minchener, 2005). The carbon conversion
bubbling fluidized bed, and the circulating fluidized bed reactor. is higher due to the high operating temperature (Ruiz et al., 2013);
The bubbling type consists of one single reaction chamber and however, the gasifier is restricted with respect to the material of
fluidization control may be a problem. A major disadvantage of construction. It is vital to implement an effective cooling technique
this reactor type is that O2 is trapped in the bubbles, and com- for the produced syngas for sulfur removal and gas handling im-
bustion may take place anywhere in the gasifier, which reduces provement. A detailed comparison of these three types of gasifiers
the efficiency of the gasifier (Ruiz et al., 2013). In the circulating is presented elsewhere (McKendry, 2002).
design, solids move in a cycle between the reactor vessel and a The heat transfer mechanism is usually ignored since the en-
cyclone separator, where ash is removed and the remaining char ergy required by the endothermic gasification reactions in all
20 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

gasifiers is directly provided by the reaction gases produced after ash and react with other ash components, such as silica, to form
partial combustion, which is one reason the heating rate in in- low melting point silicates (Kosminski et al., 2006a). The ash
dustrial gasifiers is much faster than in any laboratory kinetic melting behavior may be due to calcium silicate compounds and
study using indirect heating (Tremel et al., 2013). additives rich in the silicate mineral, quartz, which lower the
melting temperature (Coda et al., 2007).
2.3.3. Solid by-product disposal Potassium is beneficial in lowering the melting point of silica;
There are two operational types of flow gasifiers: slagging and however, it has disadvantages at certain slagging temperatures.
non-slagging. In slagging types, the operating temperature is Since alkali compounds have high vapor pressures, they tend to
above the melting temperature of the contained ash; and, the suffer losses due to sublimation under high gasifier temperatures
molten ash flows down the wall of the gasifier and forms a solid (Jones et al., 2006; Nzihou et al., 2013). If the slagging temperature
slag layer. The slag layer protects the wall material from further can be maintained below 1260 °C, potassium oxide (K2O) losses
corrosion. In non-slagging types, the operating temperature is can be minimized to about 30% with the addition of rice straw ash
below the melting temperature of the produced ash; therefore, no (Thy et al., 2005). A gasifying agent has been found to influence
solid slag layer is formed. The non-slagging gasifier is more sui- the melting point of Na2CO3. Steam is effective in lowering the
table for raw materials with lower ash contents (Tremel et al., melting temperature of Na2CO3 and, consequently, the formation
2013). temperature of liquid sodium disilicate (Kosminski et al., 2006a,
This classification of flow gasifiers is well known for EFRs. Fixed 2006b). In a steam atmosphere, Na2CO3 melts at a lower tem-
and fluidized bed reactors do not operate at high temperatures; perature than its standard temperature of 851 °C (Kosminski et al.,
therefore, only non-slagging reactors are considered. An exception 2006b).
is the plasma gasification reactor that uses fixed or fluidized The selection of the right gasifier type is ultimately linked to
plasma gasification (Materazzi et al. 2014; Ruj and Ghosh, 2014). the mineral composition of the feedstock; extrapolation from
The plasma gasification reactor is also the best example of a high-grade coal to low-grade coal or biomass is inconvenient.
slagging reactor, because plasma torches can easily melt any bio- Moreover, slagging is usually ignored in kinetic modeling, and the
mass or coal slag (Rutberg et al., 2011). assumption of constant alkali content may not be accurate at high
The selection of either a slagging or non-slagging reactor is not temperatures.
an easy task; however, it is important to distinguish whether there
will be slag formation under specific conditions or not. This aspect
is emphasized in this review, as many authors mention the use of a 3. Variables affecting char reactivity
catalyst and its regeneration; however, the recovery of a catalyst is
impossible under slagging conditions. Temperature and pressure are the controlled operation vari-
A summary of slagging and non-slagging characteristics is ables during gasification and are directly associated with re-
presented in Table 1. The two most important parameters in the activity. The gasification rate increases at higher temperatures,
prediction of slag formation are temperature and mineral com- regardless of the gasifying agent used (Hernandez et al., 2012a).
position, because they determine the slag flow parameters, such as During pyrolysis, temperature and heating rates significantly in-
viscosity. Slag must remain in the fluid phase throughout the ga- fluence char gasification reactivity more than pressure (Kajitani
sifier and be removed continuously at the bottom for proper op- et al., 2001). However, increasing these variables decreases the
eration of a slagging reactor. Various sources suggest different slag thermal efficiency of the process and can be economically un-
viscosities varying from 8–15 Pa.s (Coda et al., 2007) to 1–10 Pa s feasible. Char surface area and total mineral content play im-
(Wang et al., 2005). Low viscosity increases slag velocity; there- portant roles in gasifier selection, since they are not constant
fore, the insulating layer that protects the wall gasifier may be lost properties. The transformation of the feedstock to char is a critical
(Schrober et al., 1985). However, the slag viscosity must not exceed factor, with major differences in the reported kinetic results be-
25 Pa s (Duchesne et al., 2012; Kong et al., 2012; Schrober et al., tween laboratory studies and industrial gasifiers.
1985) for a successful slagging operation.
The composition of the ash greatly influences its melting 3.1. Reaction temperature
temperature. Alkali metals and carbonates, such as sodium and
potassium carbonates (Na2CO3 and K2CO3), are contained in the The choice of the reaction temperature to produce syngas can
be viewed from two perspectives: energy efficiency and fuel uti-
Table 1 lization. If energy efficiency is the priority, a relatively low tem-
Summary of slagging and non-slagging characteristics. perature and minimum oxygen input should be used (Qin et al.,
2011). However, an emphasis on fuel utilization may require a high
Characteristics
temperature (about 1350 °C) and a suitable excess air ratio (ap-
Slagging ■ Ash melts at a high temperature proximately 0.35) to produce desirable yields of H2 and CO and to
■ Slag protects the gasifier wall from corrosion lower soot and tar yields (Qin et al., 2011). The optimal tempera-
■ Operating temperature is greater than ash melting ture for gasification for high carbon conversion and gas yield is at
temperature least 900 °C (Zhao et al., 2009). A wide range of temperatures
■ Slag temperature  1300 to 1500 °C
 Slag viscosity ■ Must be less than 25 Pa s
should be considered in determination of the effect on produced
■ If too low, slag velocity increases and insulation may be lost syngas (as outlined below), the reaction rate, residence time, and
■ If too high, slag may accumulate on the walls surface area.
 Ash composition ■ Alkali compounds, like sodium carbonate (Na2CO3), lower At temperatures below 1000 °C, CO yield decreases and H2
the melting point of silica
yield increases with increasing temperature (Chhiti et al., 2011;
■ Lime (CaO) also lowers the ash melting temperature
■ Potassium, particularly potassium oxide (K2O), may be lost Zhao et al., 2009; Zhao et al. 2012). This occurrence indicates that
through evaporation at high slagging temperature water–gas shift (WGS) reactions are more dominant than the
Non-slagging ■ Reactor walls are kept free from slag Boudouard reaction (Chhiti et al., 2011). For air-steam gasification,
■ Operating temperature is less than ash melting the H2/CO ratio increases from 0.2 to 1 with increases in tem-
temperature
■ Operates at 800–900 °C
perature, specifically from 750 to 1150 °C (Hernandez et al.,
2012b). Steam enhances WGS and steam-reforming reactions
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 21

(Hernandez et al., 2012b; Senapati and Behera, 2012). In contrast, 2012; Hernandez et al., 2012b; Kajitani et al., 2001; Kim et al.,
the CO/CO2 ratio decreases from 1.2 to 1.1 when the temperature 2014, Qin et al., 2011; Senapati and Behera, 2012; Tremel and
decreases from 1150 °C to 750 °C (Hernandez et al., 2012b). The Spliethoff, 2013; Zhao et al., 2009; Zhao et al., 2012). The effects of
balance between CO-producing (Boudouard, steam reforming) and total pressure (Fermoso et al., 2010; Senapati and Behera, 2012;
CO-consuming (WGS) reactions is critical (Hernandez et al., Tremel and Spliethoff, 2013; Wall et al., 2002) have shown inverse
2012b). effects between thermodynamics (maximum conversion) and ki-
The reaction rate is slower at low temperatures and takes netics (Wall et al., 2002). Boudouard and steam gasification reac-
longer to achieve high conversion (Tremel et al. 2011). Tar and tions have a net increase in gas molecules (Eqs. (1) and (2), re-
volatile compounds that condense at low temperatures may also spectively), and an increase in total pressure decreases the yield of
be produced, resulting in wall deposits and clogging of pipelines gas products. However, an increase in total pressure increases
(Hernandez et al., 2012a; Zhou et al., 2009). concentration and, theoretically, should increase gasification rates,
At temperatures above 1000 °C, both CO and H2 yields increase if the final conversion is lower than the equilibrium conversion.
with increasing temperature (Qin et al., 2011; Tremel et al., 2012; For example, Fermoso et al. (2010) reported that atmospheric ga-
Yoon et al., 2009), due to enhanced Boudouard, steam reforming sification produces higher H2 and CO yields with better cold gas
and WGS reactions. CH4 decreases due to its consumption in the efficiency (CGE) than gasification at 15 atm, indicating that the
steam-reforming reaction (Qin et al., 2011; Tremel et al., 2012; syngas yield was limited by the thermodynamic equilibrium.
Yoon et al., 2009). For air-steam gasification, the H2/CO ratio in- Table 3 shows a summary of the effects of oxygen partial
creases from 0.6 to 1 when the temperature increases from 1000 pressure and total gasification pressure as presented by Wall et al.
to 1200 °C and then remains constant up to 1350 °C (Chhiti et al., (2002). In general, it is accepted that the effects of pressure are not
2011). Tars decrease at high temperatures, due to thermal cracking significant compared with other variables, such as temperature
and steam reforming (Hernandez et al., 2012a). Above 1000 °C, and mineral composition; however, other technical problems re-
cold gas efficiency is improved (Zhou et al., 2009), and higher lated to slag formation are affected by total pressure. Some mi-
reaction rates and carbon conversion have been observed (Qian neral components that melt at low temperatures are transformed
et al., 2013). under high pressure into minerals with high melting points, such
At very high temperatures (1400–1500 °C), both H2 and CO
as mullite and sanidine. On the other hand, minerals with sig-
yields decrease. In this temperature range, the reaction shifts to
nificant contents of iron reduce the melting point of the ash (Jing
the combustion region from the gasification region (Yoon et al.,
et al., 2013).
2009). Pore size is reduced and particles fuse together due to the
diffusion of atoms to the char surface at very high temperatures,
which is known as sintering (Jones et al., 2006). At very high 3.3. Pyrolysis and char surface area development
temperatures, less syngas is produced as a result of the lower
surface area, even at high residence times (Tremel et al. 2012). The way that the char is usually produced is not considered
Table 2 presents a summary of the effects of temperature on during experimental kinetic studies (Ahmed and Gupta, 2011;
syngas production, considering the corresponding reactions and Fermoso et al., 2008; Fermoso et al. 2010; Gil et al., 2012; Her-
reactivity. nandez et al., 2012a; Hernandez et al., 2012b; Kim et al., 2014;
McKendry, 2002; Ochoa et al., 2001; Rutberg et al., 2011). How-
3.2. Total pressure and gasifying agent partial pressure ever, the experimental procedure can affect the char reactivity
during gasification, because the most common gasification pro-
Several works have reported changes in the partial pressure of cedures follow the proximate analysis protocol using CO2 or steam
the gasifying agent by varying the proportion of gasifying agent in the last step, instead of air or oxygen as indicated by the ASTM
with respect to the inert gas used as a carrier gas (Ahmed and D5142 standard. There is no consensus and accurate information
Gupta, 2011; Chhiti et al., 2011; Fermoso et al., 2008, Gil et al., about the relationship between pyrolysis and gasification in the
literature; therefore, it is difficult to find a comparison of the
Table 2
Summary of the effects of increasing temperature on syngas production and re- Table 3
action rate. Summary of the effects of oxygen partial pressure and total gasification pressure
(Reprinted with permission from Wall et al. (2002)).
Temperature Characteristics
Coal chemical/physical Effect of pressure on the process
Low (o 1000 °C) – CO decreases, H2 increases, WGS reactions more process
dominant than Boudouard reactions
– H2/CO ratio increases for air-steam gasification due to Char combustion Rate ↑ with increasing O2 partial pressure at a
WGS and steam reforming reactions fixed total pressure
– CO/CO2 ratio decreases due to a balance between CO- Char combustion Rate first ↑ and then ↓ with increasing total
producing and CO-consuming reactions pressure at fixed O2 mole fraction
– Slower reaction rate CO/CO2 ratio Rate ↓ with increasing O2 partial pressure
– Longer residence time to achieve high conversion Char temperature ↑with increasing O2 partial pressure at a fixed
– Tars may be produced total pressure
High ( 4 1000 °C) – CO and H2 increase due to Boudouard, steam reform- Char gasification Rate ↑ with increasing reactant gas pressure
ing and WGS reactions Pyrolysis volatile yield ↓with increasing total pressure
– CH4 decreases due to steam reforming Char reactivity ↓with increasing pyrolysis pressure
– H2/CO ratio increases for air-steam gasification Swelling property First ↑ and then ↓ with increasing pyrolysis
– Tars decrease due to thermal cracking and steam pressure
reforming Average char porosity ↑with increasing total pyrolysis pressure
– Improved cold gas efficiency (CGE) Initial char surface area ↓with increasing total pyrolysis pressure
– Higher reaction rates and carbon conversion Heat transfer No effect of total pressure ( o20 atm) on gas
Very high (  1500 °C) – H2 and CO decrease due to sintering conductivity
– Reaction shifts to combustion from gasification region Homogeneous reaction Rate ↑ with increasing total pressure
– Particles collapse and shrink Bulk diffusivity ↓with increasing total pressure
– Lack of surface area Knudson diffusivity No effect on total pressure
22 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

results of chars produced with different methods as presented by high-rank coals with similar ash contents (Di Blasi, 2009; Irfan
Silbermann et al. (2013). et al., 2011); however, reactivity also depends on the amount of
Senneca et al. (1997) provided an explanation for the reactivity alkali and alkaline earth metals.
differences of char produced with different holding times before Reactivity is reduced when ash-free coal is gasified, which can
the gasification and proposed a correlation between deactivation be observed by comparing the gasification of the parent coal and
of char surface and its exposure to high temperatures and pres- its respective ash-free coal (Kong et al., 2014; Kopyscinski et al.,
sures. The introduction of the thermal annealing concept to ex- 2013; Kopyscinski et al., 2014a; Kopyscinski et al., 2014b). Different
plain changes in the properties of char surfaces was very im- authors have presented alternative methods to obtain ash-free
portant; however, the theory is still incomplete, and no experi- coal (Rivolta Hernandez et al., 2012; Kong et al., 2014); however, it
mental tests to correlate surface area and reactivity reduction have is important to determine whether it is beneficial to remove the
been performed in further studies about thermal annealing. Other ash and, in a further step, add an alkali-based catalyst. It is im-
authors have found that there is no agreement between the ex- portant to understand the effect of the mineral content as a cat-
perimental results and the expected activity of a specific char if the alyst and the advantage of low-grade coals in implementing
history from coal to char is omitted (Ochoa et al., 2001; Tremel and gasification.
Spliethoff, 2013). The total mineral content is determined using the ASTM D5142
If pyrolysis is considered as a part of the whole gasification, as standard and is equivalent to the ash content determined in the
shown in Fig. 3, its residence time in industrial gasifiers is ne- proximate analysis. This general classification is useful in com-
cessarily lower than the residence time of gasification, because bustion; however, it is incomplete for gasification, since inorganic
pyrolysis is much faster than gasification, irrespective of the ga- species act as a catalyst for the reaction. High ash content mate-
sifying agent. However, when pyrolysis and gasification are per- rials are definitely less desirable for combustion, but their gasifi-
formed in different reactors or intentionally separated, the pyr- cation reactions are much faster than those feedstocks containing
olysis residence time is significantly increased, and the produced low ash (Di Blasi, 2009; Irfan et al., 2011); therefore, co-feeding
char is less reactive (Ochoa et al., 2001; Senneca et al., 1997; with low-grade coal or biomass is an alternative to improve the
Senneca and Cortese, 2014). Unfortunately, recent works do not overall gasification process.
consider this important aspect of gasification, since the total time The mineral composition determined by the ash analysis is
that the sample is held at high temperatures is not controlled and usually performed using the ASTM D3682 standard, with ash ob-
not negligible with respect to the total residence time. Therefore, tained at 750 °C. This analysis is usually accurate for gasification at
kinetic studies do not model the exact industrial scale behavior. low temperatures; however, potassium can evaporate (Kopyscinski
Regardless of the solid feedstock (e.g., coal, biomass, petcoke), et al. 2013) or produce low melting point slag in the reaction
the controlled gasification variable associated with pyrolysis is the temperature range (Tafur-Marinos et al., 2014). Therefore, extra-
heating rate. As the heating rate increases, the char becomes more polation of results at high temperatures must be analyzed in detail
reactive (Di Blasi, 2009; Tremel et al., 2011; Tremel et al. 2012), to avoid overestimation of the active mineral amounts. On the
which is a well-known limitation of many experimental setups, other hand, the comparison of kinetic parameters at low tem-
such as thermogravimetric analyzers (TGAs). To overcome this peratures is not easy if the reaction is not chemically controlled
situation, some researchers have used a drop furnace reactor and free of thermodynamic limitations; for example, a comparison
(Kajitani et al., 2001; Umemoto et al., 2013), an EFR (Tremel et al., of activation energies (EA) between CO2 gasification (Kopyscinski
2011; Tremel et al., 2012) and other reactor setups (Umeki et al., et al., 2013) and steam gasification (Kopyscinski et al., 2014b) using
2012a) to produce char, subsequently performing char gasification the same ash-free coals yields unpredicted results, such as EA of
in batch reactors with TGA or gas product analyses. CO2 o EA of steam.
An unusual influence of the heating rate on the char surface There are different types of raw materials that can be used as
was reported by Seo et al. (2011); they showed that the pore feedstock for gasification, and their ash analysis is performed as
surface area exhibits a maximum for a particular heating rate. This described for coal. The most common feedstock materials are coal,
phenomenon was quantified using the Brunauer–Emmett–Teller biomass, and petcoke, which can be used as raw material for ga-
(BET) method at different heating rates; however, it may be a sification if its low reactivity can be enhanced, thereby reducing
consequence of heat transfer limitations. Control of the heating problems related to petcoke's disposal and improving the energy
rate is only possible if gasification and pyrolysis occur in different balance in oil sands upgrading (Duchesne et al., 2012; Lee et al.,
reactors, which is not an economical option; hence, there is no 2012). Biomass is beneficial to the environment, because it is CO2
reason to reduce the heating rate. neutral when using straw and other non-edible materials (Chhiti
The EFR has been used to produce char and minimize the et al., 2011). Biomass has high reactivity, due to its high contents of
changes on the char surface at the laboratory and bench scales volatiles and alkali compared with coal (Chhiti et al., 2011; Qin
(Kajitani et al., 2006; Tremel and Spliethoff, 2013); however, re- et al., 2011; Roll and Hedden, 1994).
spective gasification kinetic studies were performed in a different The key factor is the balance between the price of the raw
reactor, i.e., a high-pressure TGA, which is a semi-batch reactor. material and the reactivity of the feedstock. The reactivity is di-
Complete kinetic studies of direct gasification without separating rectly related to the surface area and alkali content Hattingh et al.,
pyrolysis from gasification using an EFR have not been conducted, (2011). Alkali and alkaline earth metals undergo different catalytic
as this technology is only operational in power plants fueled by activity to promote CO2 gasification (Mitsuoka et al., 2011; Umeki
coal (IGCC). It is now evident that the understanding of the reactor et al., 2012a) and steam gasification (Coetzee et al., 2013; Sharma
configuration and the sequence of reactions taking place are cri- et al., 2008), with an effectiveness order of potassium
tical to the performance optimization of industrial gasifiers. (K) 4sodium (Na)4 calcium (Ca) 4 magnesium (Mg) (Huang et al.,
2009). Wang et al. (2009) reported a reduction in CH4 formation
3.4. Mineral content and composition when potassium carbonate (K2CO3) was added to the feedstock.

In some circumstances, low-rank coals (rank is associated with 4. Kinetic studies


morphological changes from peat to anthracite) are considered the
same as low-grade coals (grade is associated with high ash con- This section provides an explanation for the differences be-
tent). The reactivity of low-rank coals is generally higher than tween single-step kinetic models and multi-steps kinetic models,
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 23

such as the Langmuir–Hinshelwood models. The reaction me- corrosive under reaction conditions (Marrone and Hong, 2009;
chanism for gas–solid reactions indicates that multiple steps Tomlinson and Cory, 1989); therefore, it is not possible to perform
should be considered; however, an exact reaction sequence is steam gasification using the conventional setup used for CO2 ga-
implied or a reaction mechanism is assumed in order to determine sification. Different authors have presented specific setups, such as
kinetic parameters. Therefore, a single-step kinetic model can be batch reactors (for the analysis of the product gas) (Ahmed and
used if the reaction takes place in the temperature range where Gupta, 2011; Nipattummakul et al., 2010; Su and Perlmutter, 1985;
the reaction is chemically controlled (Khalil et al., 2009). Umeki et al., 2012b), TGA (Coetzee et al., 2013; Fermoso et al.,
The selection of a particular chemical reaction model depends 2011), and flow reactors (Hernandez et al., 2012b; Detournay et al.,
on the observed trend of the experimental information since ex- 2011); however, experimental procedures are not standardized,
perimental data are fitted to the selected model by square least but are similar to those used for CO2 gasification.
regression. It is important to avoid misinterpretation of the kinetic Kinetic modeling of steam gasification is usually reported using
results due to competitive steps in catalytic gasification integrated single-step reaction models, similar to those used for CO2 gasifi-
with the effects of mass and heat transfer. The number of refer- cation (Ren et al., 2013; Silbermann et al., 2013; Zou et al., 2006).
ences related to CO2 gasification is significantly higher than steam In the literature, intrinsic kinetic studies have revealed that the
gasification, due to the technological challenges when dealing particle size and gas flow rate limit the overall reaction rate. In
with steam at high temperatures; however, there are no consistent contrast, the amount of the sample is usually disregarded, and
criteria in the literature, and results for the kinetic parameters vary there are studies that have used samples of less than 15 mg (Fer-
among authors (Di Blasi, 2009). moso et al., 2011; Silbermann et al., 2013; Zou et al., 2006) and
others with samples of more than 100 mg (Ahmed and Gupta,
4.1. CO2 gasification 2011; Nipattummakul et al., 2010; Su and Perlmutter, 1985; Umeki
et al., 2012b), without considering the mass transfer effects.
Different studies related to CO2 gasification have been reported Steam gasification is much faster than CO2 gasification when
using TGA (Coetzee et al., 2013; Gil et al., 2012; Huang et al., 2009; compared under the same experimental conditions (Ahmed and
Kajitani et al., 2001, 2006; Khalil et al., 2009; Kopyscinski et al., Gupta, 2011; Marquez-Montesinos et al., 2002; Ren et al., 2013);
2013; Kopyscinski et al., 2014a; Ochoa et al., 2001; Senneca et al., and, steam has a lower EA with respect to the CO2 gasifying agent
1997; Silbermann et al., 2013; Zhu et al., 2008) or other methods to (Marquez-Montesinos et al., 2002; Ren et al., 2013). Fig. 6 shows
determine the progress of the reaction, such as gas analysis using the comparison between CO2 and steam gasification rates of
gas chromatography (Ahmed and Gupta., 2011; Hernandez et al., woodchips char at different gasifying agent partial pressures, de-
2012c; Kajitani et al., 2006; Wang et al., 2009; Zhao et al., 2009, monstrating that steam gasification is much faster than CO2
2012). The main reason for the use of TGA is the continuous gasification.
readings and more accurate information when data are recorded
along the reaction progress (Wall et al., 2002). However, one dis- 4.3. Assumptions and simplifications within the literature
advantage is that selectivity of competitive reactions cannot be
analyzed; only global kinetics may be estimated. Thus, TGA ex- A challenge in gasification research is that many of the as-
perimentation for CO2, with its lower uncertainty, is utilized more sumptions in kinetic modeling have not been clearly presented,
often than other product analysis techniques. and the understanding of the implicit considerations of every
The most common kinetic methods are related to isothermal single model is an onerous task. Three implicit assumptions must
batch experiments, since only one experimental run is required to be reviewed in order to improve kinetic analysis: mass transfer
obtain the rate constant and reaction order for a particular com- limitation, char synthesis, and existence of a maximum reaction
bination of temperature and pressure values (Fogler, 2006). Single- rate.
step models are an extension of coal combustion to gasification
and use the same assumptions about changes on the char surface 4.3.1. Mass transfer limitation
area with respect to the grade of conversion (Loewenberg and Below 1000 °C, the chemical reaction is the controlling step;
Levendis, 1991; Su and Perlmutter., 1985). and, it is assumed that mass transfer does not have a significant
There is an overlap between pyrolysis and gasification when effect (Ahmed and Gupta, 2011; Khalil et al., 2009). This assump-
they are sequentially carried out in the same reactor, because the tion must be revised, because it can affect results and is only re-
sample should be heated from ambient conditions to the reaction lated to intrapore diffusion (Fogler, 2006). However, particle size
temperature; however, this overlap is not significant, and different information has been omitted, and chemical control is considered
experimental methods can be compared by changing the time the for experiments with samples of more than 1 g (Ahmed and
sample is held at the reaction temperature before the inert gas is Gupta, 2011; Gil et al., 2012). Interparticle diffusion also plays an
replaced by the gasifying agent (Silbermann et al., 2013). Alter- important role that cannot be overlooked (Mandapati et al., 2012)
natively, the experiment can be evaluated in two stages: char and is dependent on the internal geometrical configuration of the
synthesis and char gasification, e.g., producing char in an EFR and reactor.
performing kinetic studies with TGA (Kajitani et al., 2006). An
important flaw is that any thermal treatment at high temperature 4.3.2. Char synthesis
induces changes in the char surface area (Ochoa et al., 2001; The char synthesis methodology is usually ignored or at least
Senneca et al., 1997; Senneca and Cortese, 2014). not mentioned in technical reports or related literature. Most of
At the temperature range between 650 °C and 750 °C, the the kinetic studies performed in an EFR have been designed to
Boudouard reaction is thermodynamically limited (Smith et al., reduce the effect of a low heating rate (Chhiti et al., 2011; Her-
2005); and, a comparison of EA at different temperature ranges is nandez et al., 2012b, 2012c; Kajitani et al., 2006; Senapati and
not appropriate. Behera, 2012; Tremel et al., 2012; Zhao et al., 2009, 2012), with
further analyses at low temperatures using TGA (Kajitani et al.,
4.2. Steam gasification 2006; Tremel et al., 2012); however, cooling down and reheating
the sample may induce changes on the char surface that cannot be
When using the same feedstock, the steam gasification rate is easily detected.
significantly slower than the combustion rate. Steam is highly Comparisons of different experimental works are difficult,
24 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

(b)
(a) 1.2 bars
1.5 bars

(c) (d)
0.9 bars 0.5 bars

Fig. 6. Comparison of CO2 and steam gasification rates at different gasifying agent partial pressures: (a) 1.5 bars, (b) 1.2 bars, (c) 0.9 bars, (d) 0.5 bars (Modified with
permission from Ahmed and Gupta (2011)).

because the way that char is produced is not well explained and controlling step and mass transfer does not have a significant ef-
not all elements are considered (Di Blasi, 2009; Ochoa et al., 2001). fect (Khalil et al., 2009; Tremel and Spliethoff, 2013). Brief de-
The main assumption in most kinetic studies is that the pyrolysis scriptions of the most common single-step chemical reaction ki-
step yields a unique char (using the same temperature, sweeping netic models are as follows.
gas, and heating rate). It is also assumed that char surface char-
acteristics are not affected by chemical treatments; however, Kong  The volumetric model (VM) is the simplest model describing a
et al. (2014) have presented that the coal surface area is affected by gas–solid reaction. It assumes that the reaction takes place
acid treatment, and reactivity of ash-free coal is extremely low and uniformly within the volume of the particle. The volumetric
almost the same regardless of the raw feedstock. model is given by:

4.3.3. Existence of a maximum reaction rate


For most related studies, it is implicitly accepted that the ex-
istence of a maximum reaction rate at the beginning of the gasi- dX
fication is a consequence of changes on the char surface. One ex- = k VM (1 − X )
dt (6)
ception is Silbermann et al. (2013). This assumption is emphasized
when kinetic models are compared either for CO2 or steam gasi-
fication to fit the experimental results (Ahmed and Gupta, 2011;
Bhatia and Vartak, 1996; Bhatia and Perlmutter, 1980; Fermoso  The shrinking core model (SCM) assumes that the reaction oc-
et al., 2008, 2011; Gil et al., 2012; Kajitani et al., 2001, 2006; Ko- curs only on the surface of a shrinking carbon core. The main
pyscinski et al., 2013; Mandapati et al., 2012; Malekshahian and assumption is that the reaction moves from the surface towards
Hill, 2011; Ochoa et al., 2001; Singer and Ghoniem, 2011; Su and the interior of the particle. The model is described by:
Perlmutter, 1985; Umemoto et al., 2013; Zou et al., 2006).
The char synthesis and maximum reaction rate assumptions dX
are implicit, and there is a lack of consensus in the scientific = kSCM (1−X )2/3
dt (7)
community on the differences of reported kinetic parameters (Di
Blasi, 2009).  The integrated core model (ICM) considers n as a parameter
related to the reaction order. This is also known as the power-
4.4. Chemical reaction kinetic models for char gasification law model, and it is one of the most widely used kinetic models
for gas–solid reactions. The reaction order can be associated
Single-step chemical reaction models are the simplest expres- with the stoichiometric coefficient for elementary reactions.
sions representing the kinetic behavior of gas–solid reactions if However, for heterogeneous reactions, it can also be interpreted
adsorption and desorption are not considered or if they can be as a form factor for different particle geometries, if the diffusion
assumed as not limiting factors for the overall reaction. Single-step of the gasifying agent controls the overall gasification rate. For
reaction models can be considered a good representation of the spheres, n ¼2/3; cylinders, n ¼1/2; and flat plates, n ¼0.
gas–solid reaction if the temperature range of the reaction takes
place below 1000 °C, where the chemical reaction is the
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 25

dX RPMs. Modeling improvement is commonly attached to an in-


= kIM (1−X )n
dt (8) crease in the number of fitting parameters; however, these im-
provements do not necessarily translate to a better understanding
of the reaction mechanism.
 VM, SCM and ICM are unable to describe a maximum value of
 The normal distribution function (NDF) is able to describe the
reaction rate. This disadvantage led different authors to con-
gasification rate, even if the maximum is at X¼0 (Zou et al.,
sider the random pore model (RPM) as the best kinetic model
2006). Parameters are estimated by using nonlinear regression
(Ahmed and Gupta, 2011; Bhatia and Perlmutter, 1980; Bhatia
instead of a determined conversion assumption, which makes
and Vartak, 1996; Fermoso et al., 2008; Fermoso et al., 2011; Gil
this model easier to implement. The reaction rate can be
et al., 2012; Kajitani et al., 2001; Kajitani et al., 2006; Ko-
expressed as a function of the intrinsic rate of reaction multi-
pyscinski et al., 2013; Mandapati et al., 2012; Malekshahian and
plied by a normal probability density:
Hill, 2011; Ochoa et al., 2001; Singer and Ghoniem, 2011; Su and
Perlmutter, 1985; Umemoto et al., 2013). The RPM considers two
competing effects of structural changes during the reaction.
There is a growth of accessible pores in the initial state of ⎛ − X − X 2⎞
r=
dX
=kNDF⋅exp ⎜⎜
( m)

gasification and the coalescence or overlapping of neighboring
dt α ⎟
pores’ surfaces, which reduces the area available for reaction ⎝ ⎠ (13)
(Bhatia and Perlmutter, 1980; Bhatia and Vartak, 1996). The where α is a fitting parameter associated with the width ( w ) of the
overall reaction rate is given by: r vs. X curve. At the reaction rate of r = rm/2, Zou et al. (2006)
consider α = 2w 2.
For all of the aforementioned models, the rate constant (kmodel)
dX is considered as a parameter at isothermal conditions. This para-
= kRPM (1 − X ) 1−ψ ln (1−X )
dt (9) meter is a function of the reaction temperature as stated by the
Arrhenius equation. In addition, the effect of the total pressure is
where ψ represents a parameter that describes the internal
included in this term, but it can also be an independent function if
structure of the non-converted particle char. The definition of ψ
the total pressure is not constant along the reaction progress.
is given by:
4ΠL 0 (1−ε0 )
ψ= 4.5. Langmuir–Hinshelwood kinetic models
S02 (10)

where S0 is the pore surface area per volume, L 0 is the pore length Langmuir–Hinshelwood (LH) kinetic models were developed
per volume, ε0 is the solid porosity, and ψ is a structural parameter based on adsorption/desorption theories, and they are widely
calculated using a reduced quantity, such as t/t0.5 (Malekshahian adopted if there is competition between species (Umemoto et al.,
and Hill, 2011): 2013). If the reaction temperature increases, the controlling step is
intrapore diffusion. Even when there is no theoretical reason to
t 1−ψ ln (1−X ) −1
= use LH models under these conditions, these models may fit the
t0.5 1−ψ ln ( 1−X 0.5 )−1 (11) experimental results better due to a higher number of regression
parameters. For this reason, some authors consider the extra-
The method described by Eq. (11) to estimate ψ is not re-
polation of LH kinetic data at higher temperatures, as proposed by
presentative of the reaction mechanism if r versus X is not linear or
Kajitani et al. (2006) for CO2 coal gasification.
shows a maximum. A different approach has been presented
Using a simplified version as presented by Umemoto et al.
elsewhere (Bhatia and Perlmutter, 1980 and Zou et al., 2006),
(2013), the elementary reactions are:
where ψ is determined with known experimental information
when a maximum reaction rate ( Xm ) is observed: i1

ψ =2 ( 1 − ψ ln (1 − Xm ) ) (12) CO2 ← ( CO2 )

When a maximum is not observed, rm is given at Xm ¼0, and i2 (14)


the value of ψ in Eq. (12) is equal to 2, which is the minimum value
of this parameter if a maximum really exists (Bhatia and Perl- i3
mutter, 1980). This result is very important since many authors →
CO ← (CO)
using this model are reporting values of ψ lower than the stated
minimum value (Ahmed and Gupta, 2011; Fermoso et al., 2011; i4 (15)
Kajitani et al., 2006; Kopyscinski et al., 2013; Mandapati et al.,
2012; Umemoto et al., 2013), which contradicts its theoretical i5
C + ( CO2 ) → 2CO (16)
nature. In fact, if the model is considered with a maximum rate, ψ
should be higher than 2 (Gomez and Mahinpey, 2015b; Silber-
mann et al., 2013; Zou et al., 2006). i6

The RPM has been widely accepted due to its nonlinear de- H2 O ← ( H2 O)
pendence on char surface, which can predict a maximum reaction
i7 (17)
rate as observed experimentally since Bhatia and Perlmutter pro-
posed it in 1980, with a modification (Bhatia and Vartak, 1996) for
gas–solid reactions. Different modifications of the original model i8
and their applications to fit experimental data have been reported; →
H2 ← ( H2 )
for example, some of the most recent works present extended
i9 (18)
(Kopyscinski et al., 2013) and adaptive (Singer and Ghoniem, 2011)
26 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

i10
C + H2 O → CO+H2 (19)

Without considering the reduction in char area, LH models for


CO2 gasification and steam gasification are expressed in Eqs. (20)
and (21), respectively:
k11P CO2
r=
1+k12 P CO2 + k13 PCO (20)

k21P H2O
r=
1+k22 P H2O + k23 P H2 (21)

The complete expression, including the change on char surface,


is obtained using two factors: one corresponds to an LH model,
and the other corresponds to the intrinsic kinetic model without
the rate constant. For example, when the RPM is considered, Eq.
(22) shows a simplified model proposed by Kajitani et al. (2006),
obtained by substitution of Eq. (20) with the rate constant from Eq.
(9):
k11P CO2
r= (1 − X ) 1−ψ ln (1−X )
1+k12 P CO2 + k13 PCO (22)

Adsorption and desorption are less significant at high tem-


peratures, since the heat of adsorption is at the same magnitude as
the heat of condensation (Fogler, 2006); therefore, many of the
advantages of using LH models at high temperature lack theore-
tical support. These models present a better fit, because they have
more regression parameters; however, more parameters do not
necessarily offer a better explanation for the gasification Fig. 7. Comparison of three experimental methods on gasification rate and ex-
mechanism. istence of a maximum rate: (a) experimental method, (b) gasification rate vs.
Steam gasification occurs much faster than CO2 gasification, but conversion. Genesee coal gasification at 900 °C. (Adapted with permission from
Gomez et al. (2014)).
there is no comprehensive study in the literature that indicates the
relative conversion rates of both reactions. A few studies have
Coal Gasification: Including initial gas mixing time
compared these two reactions under similar conditions (Ahmed
1.20
and Gupta, 2011; Marquez-Montesinos et al., 2002; Ren et al.,
2013), showing that steam gasification rate is more than twice the
1.00
CO2 gasification rate, as shown in Fig. 6.
0.80
dX/dt

5. New findings 0.60

Authors commonly propose kinetic models with a higher 0.40


number of fitting parameters and a better coefficient of determi-
nation (R2) (Kopyscinski et al., 2013; Singer and Ghoniem, 2011), 0.20
but these models may not be necessarily related to the reaction
mechanism. Kinetic parameters should be theoretically in- 0.00
dependent of the kinetic model; however, there is evidence that 0.0 1.0 2.0 3.0 4.0 5.0 6.0
the EA (activation energy) estimation is correlated with the se- time [min]
lection of a particular kinetic model (Hernandez et al., 2012a, Gil BD Steam Genesee Steam BD CO2 Genesee CO2
et al., 2012; Khalil et al., 2009; Kopyscinski et al. 2013; Malek-
Fig. 8. Conversion rate vs. conversion of two Alberta coals (Boundary Dam “BD” and
shahian and Hill, 2011; Silbermann et al., 2013). This correlation Genesee) using steam and CO2 as gasifying agents at 900 °C. (Adapted with per-
contradicts the Arrhenius equation; therefore, the conventional mission from Gomez and Mahinpey (2015c)).
approach of selecting kinetic models by comparing their R2 is no
longer valid (Gomez and Mahinpey, 2015b). This may be critical if kinetic model?
the estimated EA is different from the real EA. (4) Are all assumptions correct about the existence/absence of
Most of the experimental studies and kinetic models used for mass transfer limitations at low temperatures?
gasification have implicit assumptions. Reported kinetics cannot
be considered as intrinsic kinetics if the following questions are 5.1. Real nature of the maximum gasification rate
not sufficiently answered:
Bhatia and Perlmutter (1980) explained that when the gasifi-
(1) How does the experimental procedure influence the interpretation cation rate is plotted against conversion, the maximum rate ob-
of the reaction mechanism? served is associated with changes in char surface. The maximum
(2) What are the most significant variables affecting the gasifi- gasification rate depends only on the structural parameter ψ as
cation rate, regardless of the feedstock? presented in Eq. (12), and for the same reason, it is independent of
(3) What is the real EA if the result is affected by the selected other variables such as the gasifying agent's partial pressure.
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 27

However, there is a deviation of the experimental results with Coal gasification rates with new experimental method
respect to their theory because the conversion to achieve the 1.80

maximum gasification rate is not constant. The maximum rate 1.60


occurs at a lower conversion when the reaction temperature (Irfan 1.40
et al., 2011; Kopyscinski et al., 2013; Mandapati et al., 2012; Mal-
1.20
ekshahian and Hill, 2011; Umeki et al., 2012a) or partial pressure of
the gasifying agent (steam or CO2) (Ahmed and Gupta, 2011) 1.00

dX/dt
decreases. 0.80
Different studies explaining this situation have been reported
0.60
in the literature, e.g., the non-uniform pore distribution condition
(Singer and Ghoniem, 2011). However, the time it takes to observe 0.40
a maximum gasification rate is constant, independent of the par- 0.20
tial pressure (Ahmed and Gupta, 2011), reaction temperature (Ren
0.00
et al., 2013), gasifying agent (Ahmed and Gupta, 2011; Gomez and 0.0 1.0 2.0 3.0 4.0 5.0 6.0
Mahinpey, 2015c), steam-to-fuel ratio (Nipattummakul et al., time [min]
2010), and addition of a catalyst (Wang et al., 2009). This is shown BD Steam Genesee Steam BD CO2 Genesee CO2
in Fig. 6, where the dotted line confirms that the time to observe a
Fig. 10. Application of the new experimental method proposed to correct gasifi-
maximum is constant at different partial pressures as presented by cation rates showing a maximum rate (Fig. 8). Genesee coal gasification at 900 °C.
Ahmed and Gupta (2011). (Adapted with permission from Gomez and Mahinpey (2015c)).
The observation of a maximum gasification rate is associated
with the experimental procedure, as presented for the first time by
extended to any gas–solid reaction when changing the reaction
Silbermann et al. (2013) and evaluated by Gomez et al. (2014) for
medium. Fig. 8 shows the comparison of two different coals from
CO2 gasification in a TGA at atmospheric pressure and coal of
Western Canada using CO2 and steam as gasifying agents with a
particle size o90 mm. This is shown in Fig. 7, which shows three
conventional experimental method (heating with nitrogen and
experimental methods (Fig. 7A) and compares their gasification
changing it for the gasifying agent at the reaction temperature),
rates as a function of the conversion (Fig. 7B). The initial time of
proving that the time to observe a maximum rate is independent
the gasification for methods 1 and 2 is when the system reaches
of the gasifying agent and type of coal and is just related to the
the reaction temperature and for method 3 is after 60 min of
residence time distribution of the gases within the reactor.
reaching the isothermal condition (the time at the reaction tem- Proof that there is no maximum gasification rate associated
perature in an inert atmosphere is denominated isothermal pyr- with the reaction mechanism is a breakthrough in gasification,
olysis). Method 1 is performed under CO2 atmosphere during because the amended maximum rate has been considered as an
pyrolysis and gasification, showing the highest conversion rate inherent part of gasification modeling even in the most recent
and without indicating a maximum rate; methods 2 and 3 present papers (Rollinson and Karmakar, 2015; Zoulalian et al., 2015). This
a maximum rate; however, method 3 shows the lowest gasifica- finding explains why kinetic studies in continuous reactors do not
tion rate. show a maximum rate, which can be attributed to the non-se-
The existence of a maximum rate has been proven to be asso- paration between pyrolysis and gasification. The most important
ciated with the dispersion of the gases into the reactor, i.e., when result is that kinetics modeling becomes simplified, with classic
the inert gas is replaced by the reaction gas (Gomez et al., 2014). kinetic models, such as power-law and first-order kinetic models,
The experimental methods presented by Gomez et al. (2014) were able to be used to fit experimental results more precisely than
based on the perturbation of a reaction system considered in complex models (Gomez et al., 2014; Silbermann et al., 2013).
steady state. This experiment revealed that the time to observe a
maximum rate is constant and related to the reactor volume and 5.2. Effect of the isothermal pyrolysis in char reactivity
volumetric flow rate. Three different tests were designed to prove
this hypothesis for CO2 (Gomez et al., 2014) and steam–coal gasi- Char mesopore surface area decreases during isothermal pyr-
fication (Gomez and Mahinpey, 2015c); in general, it can be olysis (Gomez et al., 2014), and it decreases more rapidly as the
reaction temperature increases. This decrease in surface area is
New experimental gasification method associated with the non-existence of a maximum gasification rate,
1 1000
because isothermal pyrolysis is introduced as an additional step to
900
0 separate pyrolysis and gasification during experimentation, as
0 5 10 15 20 25
-1
800 presented in Fig. 7 (Method 3). Ordinarily, the duration of the
time to replace N₂ by isothermal pyrolysis is not considered as a variable affecting char
Temperature [ºC]
Weight loss [mg]

700
-2 Non-isothermal pyrolyis Steam or CO₂
Gas: N₂ 600
reactivity, which is one of the reasons accounting for the differ-
-3 ences between laboratory studies and industrial gasifiers.
500
-4 Some authors show different reactivities of chars exposed to
Isothermal gasification
400 different isothermal pyrolysis; however, the explanation has been
-5 Gas: Steam or CO₂ 300 linked to thermal annealing (Senneca et al., 1997; Senneca and
-6
200 Cortese, 2014), even if there is no temperature gradient. The char
-7
surface area increases with increasing temperature during non-
100
isothermal pyrolysis (Tremel et al. 2011) and decreases during
-8 0
Gasification ends when weigh reading is constant isothermal pyrolysis (Gomez et al., 2014), which explains the re-
t [min] activity differences among chars synthesized from the same raw
material under different experimental conditions.
Fig. 9. New experimental gasification method considering the starting point of the
gasification when 95% of the inert gases have been replaced by the gasifying agent.
The experimental procedure proposed by Silbermann et al.
Genesee coal gasification at 900 °C. (Adapted with permission from Gomez and (2013), using CO2 for pyrolysis and gasification, minimizes the
Mahinpey (2015c)). effects of the experimental procedure in gasification kinetics
28 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

studies (Gomez et al., 2014). This particular method has demon- isothermal experimental data (Fermoso et al., 2008; Gil et al.,
strated that direct gasification (without switching gases) is suffi- 2012; Khalil et al., 2009; Kopyscinski et al., 2013; Kopyscinski et al.,
cient to evaluate CO2 gasification; however, it cannot be extended 2014b; Mandapati et al., 2012; Malekshahian and Hill, 2011; Ren
to steam gasification because pyrolysis and gasification overlap et al., 2013; Silbermann et al., 2013; Zou et al., 2006). However, the
between 450 °C and the gasification temperature. A new general assumption is that the particular kinetic model correctly re-
experimental method proposed by Gomez and Mahinpey (2015c), presents the reaction mechanism. Therefore, it is important to
as presented in Figs. 9 and 10 for CO2 gasification in a TGA at at- estimate the kinetic parameters independent of the kinetic model,
mospheric pressure (coal particle size o90 mm and heating rate but using the same experimental information used in kinetic
4100 °C/min), indicates higher gasification rates than the con- modeling. A theoretical deduction to estimate EA (De Micco et al.,
ventional method already illustrated in Fig. 8 for the same ex- 2012; Gomez and Mahinpey, 2015b) and the frequency factor
perimental setup. (Gomez and Mahinpey, 2015b), independent of the kinetic model,
The new method does not consider the time required for the has been reported. The validation of a set of kinetic models can be
concentration development of the gasifying agent; therefore, the done by comparing their kinetic parameters with those calculated
char converted between the initial gas change and the maximum from the free-model approach.
is not considered. For this reason, the gasification rates are higher The estimation method of EA proposed by De Micco et al. (2012)
and there is no induced maximum conversion rate. This procedure and Gomez and Mahinpey (2015b) are based on cumulative vari-
can be used to correct gasification rate data showing a maximum ables (residence time and conversion) instead of differential vari-
conversion rate, provided there are no mass transfer limitations. ables (conversion rates); therefore, the solution is robust, and
small variations at the beginning of the experimental procedure do
5.3. Most significant variables affecting char reactivity not affect significantly the estimation of the kinetic parameters. A
simple method to estimate the range of the frequency factor for
The char pore surface area has been identified as one of the gasification has also been presented by Gomez and Mahinpey
most important variables affecting gasification rate (Adschiri et al., (2015b). These are theoretical approaches based on a general rate
1986; Tremel et al., 2012); however, consistent trends between
law, Eq. (23), and the Arrhenius equation, Eq. (24):
reactivity and micropore char surface area at different levels of ash
content have not been reported. Another limitation is the indirect dX
r= =k (T ) f (X )
measurement of char mesopore and micropore surface areas, be- dt (23)
cause BET and Dubinin–Radushkevich (DR) characterizations are
carried out at  197 °C and 0 °C, respectively. The char micropore
surface area can be associated with its respective parent coal, and k = k o e−EA /RT (24)
a consistent trend between gasification rate and micropore area
where r is the reaction rate (expressed as conversion rate); k is the
can be observed at different levels of ash content, if the micropore
rate constant; ko and EA are the frequency factor and activation
surface area is based on carbon content instead of coal content
energy (kinetic parameters of the Arrhenius equation), respec-
(Gomez and Mahinpey, 2015a).
tively; and, f(X) is a term related to the reaction mechanism.
The alkali content affects the coal reactivity during gasification.
Combining (Eqs. (23) and 24), integrating for a particular con-
Due to the catalytic nature of alkali and alkaline earth metals, the
version and assuming the reaction mechanism does not change
gasification rate of different feedstock mixtures cannot be con-
along the conversion range yields Eq. (25) as presented elsewhere
sidered as the linear contribution of the raw species gasification
rates (Gomez and Mahinpey, 2015a), as reported in the literature (Gomez and Mahinpey, 2015b):
(Fermoso et al., 2008; Gil et al., 2012). The catalytic effect of alkali EA E
ln ⎡⎣ tX (T ) ⎤⎦ = ln [G (X )]−ln ⎡⎣ k o ⎤⎦ +
{ } =α + A
and alkaline earth metals can be quantified for a broad range of RT RT (25)
ash contents and for different feedstocks considering alkali con-
tent equivalents; it was reported that CO2 gasification is catalyzed where tX (T ) is the residence time to achieve a particular conver-
above 0.025 equivalent moles of alkali per gram of coal (Gomez sion (X) at the reaction temperature (T), and G(X) is the integrated
and Mahinpey, 2015a). form of f(X). For gasification, the term ln [G (X )] can be considered
Consideration should be given as to whether or not the ash negligible for 0.5 rXr0.8, therefore, the range of frequency factor
should be removed and, in a further step, to add a catalyst, as values can be estimated. EA can be accurately estimated at any
usually proposed in ash-free coal gasification studies This is an particular conversion over a broader conversion range
important factor to be considered since many authors are pro- (0.1r Xr0.9).
moting ash-free gasification, which is equivalent to non-catalytic Eq. 24 indicates that the EA is independent of the reaction
gasification. Thus, the co-gasification of low-grade coal and/or mechanism, since the reaction mechanism affects the intercept
biomass with high-grade coal offers more advantages than ash- and not the slope. Therefore, a kinetic model does not represent
free coal with catalyst. the real reaction mechanism if the EA estimated in the conven-
tional way (using rate constants) is different than the one esti-
5.4. Estimation of kinetic parameters independent of the kinetic mated independent of the kinetic model. Different gasification
model study cases show that VM and ICM fit the experimental results
better with consistent kinetic parameters (Gomez and Mahinpey,
The question remains as to which kinetic model should be used 2015b).
to estimate kinetic parameters and to accurately model gasifica- These new elements on the fundamentals of chemical reaction
tion kinetics after presenting new gasification concepts. In kinetic engineering are very important to determine the validity of the
studies, the experimental information is fitted to a particular kinetic studies assumptions, e.g., if the RPM really represents the
equation by least squares regression. There are different criteria to reaction mechanism of gasification. It is very common to find new
determine the best fit, with R2 being the most widely reported. papers where the novelty is the use of nonlinear kinetic models
The conventional approach to estimate EA requires the previous such as RPM (Kopyscinski et al., 2013; Rollinson and Karmakar,
determination of the rate constant at different reaction tempera- 2015; Singer and Ghoniem, 2011; Zoulalian et al., 2015); however,
tures; this process involves the selection of a kinetic model to fit deep analysis of the model validity is not presented anywhere.
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 29

5.5. Reduction of interparticle diffusion in experimental studies maximum rate is a consequence of different gasifying agent
partial pressures during the stabilization time after the inert gas
Many researchers consider the chemical reaction as the con- is replaced with the reaction gas (Gomez et al., 2014).
trolling step in kinetic studies using a single-step kinetic model at  An increase in the time that the char is held at the reaction
temperatures lower than 1000 °C (Ahmed and Gupta, 2011; Di Blasi, temperature with an inert gas (isothermal pyrolysis) decreases
2009; Fermoso et al., 2008; Gil et al., 2012; Kopyscinski et al., 2013; the gasification rate (Gomez et al. 2014). This indicates the
Nipattummakul et al., 2010; Ren et al., 2013; Silbermann et al., importance of the thermal history of the char and explains the
2013; Tremel et al., 2011). These studies assumed that there are no differences between laboratory and industrial gasifier kinetics
associated mass transfer effects, regardless of different sample sizes (Gomez et al., 2014; Senneca et al., 1997).
used in experiments, i.e., 10 mg or less (Fermoso et al., 2008, 2011;  The char micropore surface area and the catalyst content are the
Gil et al., 2012; Gomez et al., 2014; Huang et al., 2009; Khalil et al., two most important variables affecting the gasification rate.
2009; Kopyscinski et al., 2013, 2014a; Marquez-Montesinos et al., Above 0.025 equivalent moles of alkali per gram of coal, cata-
2002; Mitsuoka et al., 2011; Silbermann et al., 2013; Umemoto et al., lytic gasification with CO2 becomes dominant, therefore, a
2013), 11–100 mg (Hattingh et al., 2011; Kong et al., 2014; Malek- higher content of alkali reduces the gasification activation
shahian and Hill, 2011; Ren et al., 2013; Senneca and Cortese, 2014; energy.
Sharma et al., 2008; Tremel et al., 2011; Umeki et al., 2012a,, 2012b;  The comparison of the kinetic parameters obtained from a
Zou et al., 2006), and more than 1 g (Ahmed and Gupta, 2011; particular kinetic model with those calculated from the free-
Coetzee et al., 2013; Nipattummakul et al., 2010). Most of these model approach is a validation tool to determine whether or not
investigations considered intrinsic kinetics if the particle size was the kinetic model represents the reaction mechanism of a par-
below 90 mm (Kong et al., 2014; Malekshahian and Hill, 2011; ticular reaction.
Sharma et al., 2008; Silbermann et al., 2013); however, even at low  The main assumption during kinetic studies at low tempera-
temperatures and small particle sizes, there is interparticle diffusion tures (below 1000 °C) is that the chemical reaction is the con-
(Gomez and Mahinpey, 2015c; Mandapati et al., 2012). trolling step (Ahmed and Gupta, 2011; Di Blasi, 2009; Fermoso
Experiments comparing a commercial TGA with a modified et al., 2008; Gil et al., 2012; Kopyscinski et al., 2013; Ni-
TGA (home-built quartz crucible and reactor) showed that the pattummakul et al., 2010; Ren et al., 2013; Silbermann et al.,
sample amount affected the gasification rate. The results demon- 2013, Tremel et al., 2011); however, experiments with small
strated that the most important variable associated with inter- particle size (90 mm) and small sample amounts (10 mg) ex-
particle diffusion is the thickness of the feedstock sample (Gomez hibited interparticle diffusion limitations due to the thickness of
and Mahinpey, 2015c; Mandapati et al., 2012), which depends on the sample (Gomez and Mahinpey, 2015c; Mandapati et al.,
the geometry of the crucible or sample holder. Most mathematical 2012). It has been proven that the reported EA of steam
models consider one single particle with a particular shape, and gasification is usually underestimated (Gomez and Mahinpey,
just negligible diffusion can be if there is a monolayer of solid 2015c), as a result of mass transfer limitations, specifically
particles. This brings another question for future studies as: Are interparticle diffusion.
the effectiveness factor and Thiele module enough to determine if  The summary of findings presented in this review allows for the
a system is mass transfer limited (considering that the reaction conclusion that the most popular kinetic model used for gasi-
mechanism is unknown and there are many particles)? fication and combustion – the random pore model (RPM) – does
Steam gasification is faster than CO2 gasification; thus, mass not represent the real reaction mechanism and that the use of
transfer effects are more significant for steam gasification at the RPM may mislead researchers in their interpretation of kinetic
same temperature since the limiting step corresponds to the analysis.
slowest process. The EA estimated from conventional gasification
experiments is significantly smaller than the one calculated if in- The new findings presented in this review provide elements to
terparticle diffusion is minimized (Gomez and Mahinpey, 2015c). improve kinetic analysis, because the most used kinetic model –
This situation has been termed falsified kinetics (Fogler, 2006) and the random pore model – cannot be considered as an accurate
indicates that the assumption of intrinsic kinetics is not possible representation of the gasification mechanism. The error incurred
for steam gasification using conventional methods, even at low by the separation of pyrolysis and gasification affects most of the
reaction temperatures. A new proposed experimental procedure,
experimental works reported in the literature. Therefore, it is ad-
with negligible interparticle diffusion and without inducing a
visable to encourage other researchers to conduct experiments,
maximum gasification rate in CO2 and/or steam gasification, is an
accounting for the findings presented in this review, even stan-
alternative to conduct more precise gasification kinetic studies
(Gomez and Mahinpey, 2015c). This is presented in Fig. 9, where dardizing the experimental procedure at the laboratory scale.
consideration of the right crucible size and sample amount will While kinetics can now be studied with simpler kinetic models,
ensure existence of mono- or bi-layer solid particles within the challenges remain, such as the determination of the real CO, CO2,
crucible, thereby minimizing the mass transfer effect. and H2O concentrations on the char surface due to the water–gas
shift reaction. Small sample sizes help to reduce mass transfer
effects, but increase data dispersion, which make it almost im-
6. Conclusions possible to determine gas concentrations in real time.
Efforts in catalytic gasification, by eliminating the ash and,
The correct interpretation and modeling of the gasification subsequently, adding a specific catalyst, should be reconsidered.
mechanism cannot be achieved without incorporating new con- Ash-free feedstock (especially coal) is a great alternative for
cepts, experimental techniques, and fundamentals in chemical combustion, but it is not favorable to reduce the feedstock surface
reaction engineering:
area and remove the catalyst in an acid leaching process for the
 The maximum gasification rate observed when the conversion further addition of a more expensive catalyst. Reuse of the ash
rate is plotted against conversion is not a consequence of is a promising alternative; therefore, studies in non-conventional
changes in char surface during gasification as proposed by reactor configurations will help to increase raw feedstock gasifi-
Bhatia and Perlmutter (1980). It has been demonstrated that the cation efficiency.
30 N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31

Acknowledgments Hattingh, B.B., Everson, R.C., Neomagus, H., Bunt, J.R., 2011. Assessing the catalytic
effect of coal ash constituents on the CO2 gasification rate of high ash, South
African coal. Fuel Process. Technol. 92 (10), 2048–2054.
The authors are thankful to the Natural Sciences and En- Hernandez, J.J., Ballesteros, R., Aranda, G., 2012a. Characterization of tars from
gineering Research Council (NSERC) of Canada for funding this biomass gasification: effect of the operating conditions. Energy 50, 333–342.
study. Hernandez, J.J., Aranda, G., Barba, J., Mendoza, J.M., 2012b. Effect of steam content in
the air-steam flow on biomass entrained flow gasification. Fuel Process. Tech-
nol. 99, 43–55.
Hernandez, J.J., Aranda-Almansa, G., Bula, A., 2012c. Gasification of biomass wastes
References in an entrained flow gasifier: effect of the particle size and the residence time.
Fuel Process. Technol. 91, 681–692.
Hoffmann, B.S., Szklo, A., 2011. Integrated gasification combined cycle and carbon
Adschiri, T., Shiraha, T., Kojima, T., Furusawa, T., 1986. Prediction of CO2 gasification capture: a risky option to mitigate CO2 emissions of coal-fired power plants.
rate of char in fluidized bed gasifier. Fuel 65 (12), 1688–1693. Appl. Energy 88, 3917–3929.
Ahmed, I.I., Gupta, A.K., 2011. Kinetics of woodchips char gasification with steam Huang, Y., Yin, X., Wu, Ch, Wang, C., Xie, J., Zhou, Z., Ma, L., Li, H., 2009. Effects of
and carbon dioxide. Appl. Energy 88 (5), 1613–1619. metal catalysts on CO2 gasification reactivity of biomass char. Biotechnol. Adv.
Bhatia, S.K., Perlmutter, D.D., 1980. A random pore model for fluid  solid reactions: 27, 568–572.
I. Isothermal and kinetic control. AIChE J. 26 (3), 379–386. Irfan, M., Usman, M.R., Kusakabe, K., 2011. Coal gasification in CO2 atmosphere and
Bhatia, S.K., Vartak, B.J., 1996. Reaction of microporous solids: the discrete random its kinetics since 1948: a brief review. Energy 36, 12–40.
pore model. Carbon 34 (11), 1383–1391. Jing, N., Wang, Q., Cheng, L., Luo, Z., Cen, K., Zhang, D., 2013. Effect of temperature
Bielowicz, B., 2012. A new technological classification of low-rank coal on the basis and pressure on the mineralogical and fusion characteristics of Jincheng coal
of Polish deposits. Fuel 96, 497–510. ash in simulated combustion and gasification environments. Fuel 104, 647–655.
Bremaud, M., Fongarland, P., Anfray, J., Jallais, S., Schweich, D., Khodakov, A.Y., 2005. Jones, J.M., Darvell, L.I., Bridgeman, T.G., Pourkashanian, M., Williams, A., 2006. An
Influence of syngas composition on the transient behavior of a Fischer–Tropsch investigation of the thermal and catalytic behaviour of potassium in biomass
continuous slurry reactor. Catal. Today 106, 137–142. combustion. Proc. Combust. Inst. 31, 1955–1963.
Coda, B., Cieplik, M.J., de Wild, P.J., Kiel, J.H., 2007. Slagging behavior of wood ash Kajitani, S., Hara, S., Matsuda, H., 2001. Gasification rate analysis of coal char with a
under entrained-flow gasification conditions. Energy Fuels 21, 3644–3652. pressurized drop tube furnace. Fuel 81, 539–546.
Chhiti, Y., Salvador, S., Commandre, J., Broust, F., Couhert, C., 2011. Wood bio-oil Kajitani, S., Suzuki, N., Ashizawa, M., Hara, S., 2006. CO2 gasification rate analysis of
noncatalytic gasification: influence of temperature, dilution by an alcohol and coal char in entrained flow coal gasifier. Fuel 85 (2), 163–169.
ash content. Energy Fuels 25, 345–351. Khalil, R., Varhegyi, G., Jaschke, S., Gronli, M.G., Hustad, J., 2009. CO2 gasification of
Coetzee, S., Neomagus, H.W.J.P., Bunt, J.R., Everson, R.C., 2013. Improved reactivity of biomass chars: a kinetic study. Energy Fuels 23, 94–100.
large coal particles by K2CO3 addition during steam gasification. Fuel Process. Kihedu, J.H., Yoshiie, R., Nunome, Y., Ueki, Y., Naruse, I., 2014. Counter-flow air ga-
Technol. 114, 75–80. sification of woody biomass pellets in the auto-thermal packed bed reactor.
Collot, A., 2006. Matching gasification technologies to coal properties. Int. J. Coal Fuel 117, 1242–1247.
Geol. 65 (3), 191–212. Kim, Y., Park, J., Jung, D., Miyawaki, J., Yoon, S., Mochida, I., 2014. Low-temperature
Cousins, A., Hughes, R.W., McCalden, D.J., Lu, D.Y., Anthony, E.J., 2012. Waste clas- catalytic conversion of lignite: 1. Steam gasification using potassium carbonate
sification of clag generated in a pilot-scale entrained-flow gasifier. CSIRO supported on perovskite oxide. J. Ind. Eng. Chem. 20, 216–221.
Canmet ENERGY 2, 37–44 〈www.coalcgp-journal.org〉 (last accessed August Kong, L., Bai, J., Bai, Z., Guo, Z., Li, W., 2012. Effects of CaCO3 on slag flow properties
2014). at high temperatures. Fuel 109, 76–85.
Datta, S., Sarkar, P., Chavan, P.D., Saha, S., Sahu, G., Sinha, A.K., Saxena, V.K., 2015. Kong, Y., Kim, J., Chun, D., Lee, S., Rhim, Y., Lim, J., Choi, H., Kim, S., Yoo, J., 2014.
Agglomeration behavior of high ash Indian coals in fluidized bed gasification Comparative studies on steam gasification of ash-free coals and their original
pilot plant. Appl. Therm. Eng. 86, 222–228. raw coals. Int. J. Hydrog. Energy 39, 9212–9220.
De Micco, G., Nasjleti, A., Bohe, A.E., 2012. Kinetics of the gasification of a Rio Turbio Kopyscinski, J., Habibi, R., Mims, Ch.A., Hill, J.M., 2013. K2CO3-catalyzed CO2 gasi-
coal under different pyrolysis temperatures. Fuel 95, 537–543. fication of ash-free coal: Kinetic study. Energy Fuels 27, 4875–4883.
Dermot, J.R., 2013. A syngas network for reducing industrial carbon footprint and Kopyscinski, J., Rahman, M., Gupta, R., Mims, Ch.A., Hill, J.M., 2014a. K2CO3 catalyzed
energy use. Appl. Therm. Eng. 53, 299–304. CO2 gasification of ash-free coal. Interactions of the catalyst with carbon in N2
Detournay, M., Hemati, M., Andreux, R., 2011. Biomass steam gasification in flui- and CO2 atmosphere. Fuel 117, 1181–1189.
dized bed of inert or catalytic particles: Comparison between experimental Kopyscinski, J., Lam, J., Mims, Ch.A., Hill, J.M., 2014b. K2CO3 catalyzed steam gasi-
results and thermodynamic equilibrium predictions. Powder Technol. 208 (2), fication of ash-free coal. Studying the effect of temperature on carbon con-
558–567. version and gas production rate using a drop-down reactor. Fuel 128, 210–219.
Di Blasi, C., 2009. Combustion and gasification rates of lignocellulosic chars. Prog. Kosminski, A., Ross, D.P., Agnew, J.B., 2006a. Influence of gas environment on re-
Energy Combust. 35, 121–140. actions between sodium and silicon minerals during gasification of low-rank
Duchesne, M.A., Ilyushechkin, A.Y., Hughes, R.W., Lu, D.Y., McCalden, D.J., Macchi, A., coal. Fuel Process. Technol. 87, 953–962.
Anthony, E.J., 2012. Flow behavior of slags from coal and petroleum coke Kosminski, A., Ross, D.P., Agnew, J.B., 2006b. Reactions between sodium and silica
blends. Fuel 97, 321–328. during gasification of a low-rank coal. Fuel Process. Technol. 87, 1037–1049.
EIA (US Energy Information Administration), 2011. Natural Gas Consumption Has Lee, S.H., Yoon, S.J., Ra, H.W., Son, Y., Hong, J.C., Lee, J.G., 2012. Gasification char-
Two Peaks Each Year. 〈http://www.eia.gov/todayinenergy/detail.cfm?id ¼2050〉 acteristics of coke and mixture with coal in an entrained-flow gasifier. Energy
(last accessed August 2014). 35, 3239–3244.
EIA (US Energy Information Administration), 2013. Future Power Market Shares of Loewenberg, D.A., Levendis, W.A., 1991. Combustion behavior and kinetics of syn-
cOal, Natural Gas Generators Depend on Relative Fuel Prices. 〈http://www.eia. thetic and coal-derived chars: Comparison of theory and experiment. Combust.
gov/todayinenergy/detail.cfm?id ¼10951〉 (last accessed August 2014). Flame 84, 47–65.
Fermoso, J., Arias, B., Pevida, C., Plaza, M.G., Rubiera, F., Pis, J.J., 2008. Kinetic models Lu, T., Li, K., Zhang, R., Bi, J., 2015. Addition of ash to prevent agglomeration during
comparison for steam gasification of different nature fuel chars. J. Therm. Anal. catalytic coal gasification in a pressurized fluidized bed. Fuel Process. Technol.
Calorim. 91 (3), 779–786. 134, 414–423.
Fermoso, J., Arias, B., Gil, M.V., Plaza, M.G., Pevida, C., Pis, J.J., Rubiera, F., 2010. Co- Ma, L., Zhu, Y., Chen, H., Ding, Y., 2011. An environment friendly energy recovery
gasification of different rank coals with biomass and petroleum coke in a high- technology: Municipal solid waste gasification. In: Proceedings of the Third
pressure reactor for H2-rich gas production. Bioresour. Technol. 101, International Conference on Measuring Technology and Mechatronics Auto-
3230–3235. mation, pp. 407–409.
Fermoso, J., Gil, M.V., García, S., Pevida, C., Pis, J.J., Rubiera, F., 2011. Kinetic para- Malekshahian, M., Hill, J.M., 2011. Kinetic analysis of CO2 gasification of petroleum
meters and reactivity for the steam gasification of coal chars obtained under coke at high pressure. Energy Fuels 25, 4043–4048.
different pyrolysis temperatures and pressures. Energy Fuels 25, 3574–3580. Mandapati, R.M., Daggupati, S., Mahajani, S.M., Aghalayam, P., Sapru, R.K., Sharma,
Fogler, H.S., 2006. Elements of Chemical Reaction Engineering, Diffusion and Re- R.K., et al., 2012. Experiments and kinetic modeling for CO2 gasification of In-
action, 4th ed. Pearson Education, Inc, Boston, pp. 813–851. dian coal chars in the context of underground coal gasification. Ind. Eng. Chem.
Gil, M.V., Riaza, J., Alvarez, L., Pevida, C., Pis, J.J., Rubiera, F., 2012. Kinetic models for Res. 51, 15041–15052.
the oxy-fuel combustion of coal and coal/biomass blend chars obtained in N2 ́
Marquez-Montesinos, F., Cordero, T., Rodrıguez-Mirasol, ́
J., Rodrıguez, J.J., 2002. CO2
and CO2 atmospheres. Energy 48, 510–518. and steam gasification of a grapefruit skin char. Fuel 81 (4), 423–429.
Gomez, A., Silbermann, R., Mahinpey, N., 2014. A comprehensive experimental Marrone, P.A., Hong, G.T., 2009. Corrosion control methods in supercritical water
procedure for CO2 coal gasification: Is there really a maximum reaction rate? oxidation and gasification processes. J. Supercritic. Fluid 51, 83–103.
Appl. Energy 124, 73–81. Materazzi, M., Lettieri, P., Mazzei, L., Taylor, R., Chapman, C., 2014. Tar evolution in a
Gomez, A., Mahinpey, N., 2015a. A new model to estimate CO2 coal gasification kinetics two stage fluid bed–plasma gasification process for waste valorization. Fuel
based only on parent coal characterization properties. Appl. Energy 137, 126–133. Process. Technol. 128, 146–157.
Gomez, A., Mahinpey, N., 2015b. A new method to calculate kinetic parameters Matthews, H.D., Gillet, N.P., Stott, P.A., Zickfeld, K., 2009. The proportionality of
independent of the kinetic model: insights on CO2 and steam gasification. global warming to cumulative carbon emissions. Nature 459, 829–832.
Chem. Eng. Res. Des. 95, 346–357. McKendry, P., 2002. Energy production from biomass (Part 3): gasification tech-
Gomez, A., Mahinpey, N., 2015c. Kinetic study of coal steam and CO2 gasification: a nologies. Bioresour. Technol. 83, 55–63.
new method to reduce interparticle diffusion. Fuel 148, 160–167. Minchener, A., 2005. Coal gasification for advanced power generation. Fuel 84 (17),
N. Mahinpey, A. Gomez / Chemical Engineering Science 148 (2016) 14–31 31

2222–2235. CO2 gasification method using coal from deep unmineable seams. Ind. Eng.
Mitsuoka, K., Hayashi, S., Amano, H., Kayahara, K., Sasaoaka, E., Uddin, A., 2011. Chem. Res. 52, 14787–14797.
Gasification of woody biomass char with CO2: The catalytic effects of K and Ca Singer, S.L., Ghoniem, A.F., 2011. An adaptive random pore model for multimodal
species on char gasification reactivity. Fuel Process. Technol. 92, 26–31. pore structure evolution with application to char gasification. Energy Fuels 25,
NETL (The national energy technology laboratory), 2002. Gasification technologies: 1423–1437.
gasification markets and technologies- present and future. 〈http://www.netl. Smith, J.M., Van Ness, H.C., Abbott, M.M., 2005. Introduction to Chemical En-
doe.gov/File%20library/research/coal/energy%20systems/gasification/gasifipe gineering Thermodynamics, 7th ed. McGraw-Hill, New York.
dia/Gasification_Technologies.pdf〉 (last accessed August 2014). Su, J.L., Perlmutter, D.D., 1985. Effect of pore structure on char oxidation kinetics.
Nipattummakul, N., Ahmed, I., Kerdsuwan, S., Gupta, A.K., 2010. High temperature AIChE J. 31, 973–981.
steam gasification of wastewater sludge. Appl. Energy 87 (12), 3729–3734. Tafur-Marinos, J.A., Ginepro, M., Pastero, L., Torazzo, A., Paschetta, E., Fabbri, D.,
Nzihou, A., Stanmore, B., Sharrock, P., 2013. A review of catalysts for the gasification Zelano, V., 2014. Comparison of inorganic constituents in bottom and fly re-
of biomass char, with some reference to coal. Energy 58, 305–317. sidues from pelletised wood pyro-gasification. Fuel 119, 157–162.
Ochoa, J., Cassanello, M., Bonelli, P., Cukierman, A., 2001. CO2 Gasification of Ar- Thibault, B., 2012. Electricity From Coal: Time to Turn the Page on Canada's Dirtiest
gentinean coal chars: a kinetic characterization. Fuel Process. Technol. 74 (3), Source of Power. Pembina Institute 〈http://www.pembina.org/blog/634〉 (last
161–176. accessed July 2014).
O'Keefe, J.M.K., Bechtel, A., Christanis, K., Dai, S., DiMichele, W.A., Eble, C.F., et al., Thy, P., Jenkins, B.M., Lesher, C.E., Grundvig, S., 2005. Compositional constraints on
2013. On the fundamental difference between coal rank and coal type. Int. J. slag formation and potassium volatilization from rice straw blended wood fuel.
Coal Geol. 118, 58–87. Fuel Process. Technol. 87, 383–408.
Pian, C.C., 2002. Regenerative gasification systems operating on farm-waste and Tomlinson, L., Cory, N.J., 1989. Hydrogen emission during the steam oxidation of
bioenergy-crop feedstocks. In: Proceedings of the 37th Intersociety Energy ferritic steels: kinetics and mechanism. Corros. Sci. 29 (8), 939–965.
Conversion Engineering Conference, pp. 668–674. Tremel, A., Haselstenier, T., Kunze, C., Spliethoff, H., 2011. Experimental investiga-
Qian, F., Kong, X., Cheng, H., Du, W., Zhong, W., 2013. Development of a kinetic tion of high temperature and high pressure coal gasification. Appl. Energy 92,
model for industrial entrained flow coal gasifiers. Ind. Eng. Chem. Res. 52, 79–285.
1819–1828. Tremel, A., Haselsteiner, T., Nakonz, M., Spliethoff, H., 2012. Coal and char proper-
Qin, K., Lin, W., Jensen, P.A., Jensen, A.D., 2011. High-temperature entrained flow ties in high temperature entrained flow gasification. Energy 45, 176–182.
gasification of biomass. Fuel 93, 589–600. Tremel, A., Spliethoff, H., 2013. Gasification kinetics during entrained flow gasifi-
Ren, L., Yang, J., Gao, F., Yan, J., 2013. Laboratory study on gasification reactivity of cation – Part II: intrinsic char reaction rate and surface area development. Fuel
coals and petcokes in CO2/steam at high temperatures. Energy Fuels 27, 107, 653–661.
5054–5068. Tremel, A., Becherer, D., Fendt, S., Gaderer, M., Spliethoff, H., 2013. Performance of
Rezaiyan, J., Cheremisinoff, N.P., 2005. Gasification Technologies, A premier for entrained flow and fluidised bed biomass gasifiers on different scales. Energy
Engineer and Scientists. CRC Press, Taylor & Francis Group, Boca Raton. Convers. Manag. 69, 95–106.
Rivolta Hernandez, M., Figueroa Murcia, C., Gupta, R., De Klerk, A., 2012. Sol- Umeki, K., Moilanen, A., Gómez-Barea, A., Konttinen, J., 2012a. A model of biomass
vent coal  mineral interaction during solvent extraction of coal. Energy Fuels char gasification describing the change in catalytic activity of ash. Chem. Eng. J.
26 (11), 6834–6842. 207–208, 616–624.
Roll, H., Hedden, K., 1994. Entrained flow gasification of coarsely ground Chinese Umeki, K., Namioka, T., Yoshikawa, K., 2012b. The effect of steam on pyrolysis and
reed. Chem. Eng. Process. 33, 353–361. char reactions behavior during rice straw gasification. Fuel Process. Technol. 94
Rollinson, A.N., Karmakar, M.K., 2015. On the reactivity of various biomass species (1), 53–60.
with CO2 using a standardised methodology for fixed-bed gasification. Chem. Umemoto, S., Kajitani, S., Hara, S., 2013. Modeling of coal char gasification in co-
Eng. Sci. 128, 82–91. existence of CO2 and H2O considering sharing of active sites. Fuel 103, 14–21.
Ruiz, J., Juarez, M., Morales, M., Munoz, P., Mendivil, M., 2013. Biomass gasification Wall, T.F., Liu, G., Wu, H., Roberts, D.G., Benfell, K.E., Gupta, S., Lucas, J.A., Harris, D.J.,
for electricity generation: review of current technology barriers. Renew. Sust. 2002. The effects of pressure on coal reactions during pulverised coal com-
Energy Rev. 18, 174–183. bustion and gasification. Prog. Energy Combust. 28, 405–433.
Ruj, B., Ghosh, S., 2014. Technological aspects for thermal plasma treatment of Wang, B., Li, X., Xu, S., Paterson, N., Dugwell, D.R., Kandiyoti, R., 2005. Performance
municipal solid waste—A review. Fuel Process. Technol. 126, 298–308. of Chinese coals under conditions simulating entrained-flow gasification. En-
Rutberg, Ph.G., Bratsev, A.N., Kuznetsov, V.A., Popov, V.E., Ufimtsev, A.A., Shtengel, ergy Fuels 19 (5), 2006–2013.
S.V., 2011. On efficiency of plasma gasification of wood residues. Biomass. Wang, J., Jiang, M., Yao, Y., Zhang, Y., Cao, J., 2009. Steam gasification of coal char
Bioenergy 35, 495–504. catalyzed by K2CO3 for enhanced production of hydrogen without formation of
Schrober, H.H., Streeter, R.C., Diehl, E.K., 1985. Flow properties of low-rank coal ash methane. Fuel 88 (9), 1572–1579.
slags: implications for slagging gasification. Fuel 64 (11), 1611–1617. Warnecke, R., 2000. Gasification of biomass: comparison of fixed bed and fluidized
Schweinfurth, S.P., 2009. An introduction to coal quality. U.S. Geological Survey bed gasifier. Biomass. Bioenergy 18, 489–497.
Professional Paper 1625–F. 〈http://pubs.usgs.gov/pp/1625f/downloads/Chap Yoon, S.J., Choi, Y., Son, Y., Lee, S., Lee, J., 2009. Gasification of biodiesel by-product
terC.pdf〉 (last accessed August 2014). with air or oxygen to make syngas. Bioresour. Technol. 101, 1227–1232.
Senapati, P.K., Behera, S., 2012. Experimental investigation on an entrained flow Zhang, Q., Wu, Y., Dor, L., Yang, W., Blasiak, W., 2013. A thermodynamic analysis of
type biomass gasification system using coconut coir dust as powdery biomass solid waste gasification in the Plasma Gasification Melting process. Appl. En-
feedstock. Bioresour. Technol. 117, 99–106. ergy 112, 405–413.
Senneca, O., Russo, P., Salatino, P., Masi, S., 1997. The relevance of thermal annealing Zhao, Y., Sun, S., Tian, H., Qian, J., Su, F., Ling, F., 2009. Characteristics of rice husk
to the evolution of coal char gasification reactivity. Carbon 35 (1), 141–151. gasification in an entrained flow reactor. Bioresour. Technol. 100, 6040–6044.
Senneca, O., Cortese, L., 2014. Thermal annealing of coal at high temperature and Zhao, Y., Sun, S., Zhou, H., Sun, R., Tian, H., Luan, J., Qian, J., 2012. Experimental study
high pressure. Effects on fragmentation and on rate of combustion, gasification on sawdust air gasification in an entrained-flow reactor. Fuel Process. Technol.
and oxy-combustion. Fuel 116, 221–228. 91, 910–914.
Seo, D.K., Park, S.S., Kim, Y.T., Hwang, J., Yu, T., 2011. Study of coal pyrolysis by Zhou, J., Chen, Q., Zhao, H., Cao, X., Mei, Q., Luo, Z., Cen, K., 2009. Biomass-oxygen
thermo-gravimetric analysis (TGA) and concentration measurements of the gasification in a high-temperature entrained-flow gasifier. Biotechnol. Adv. 27,
evolved species. J. Anal. Appl. Pyrol. 92 (1), 209–216. 606–611.
Serrano, D., Sanchez-Delgado, S., Sobrino, C., Marugan-Cruz, C., 2015. Defluidization Zhu, W., Song, W., Lin, W., 2008. Catalytic gasification of char from co-pyrolysis of
and agglomeration of a fluidized bed reactor during Cynara cardunculus L. coal and biomass. Fuel Process. Technol. 89 (9), 890–896.
gasification using sepiolite as a bed material. Fuel Process. Technol. 131, Zou, J.H., Zhou, Z.J., Wang, F.C., Zhang, W., Dai, Z.H., Liu, H.F., Yu, Z.H., 2006. Mod-
338–347. eling reaction kinetics of petroleum coke gasification with CO2. Chem. Eng.
Sharma, A., Takanohashi, T., Morishita, K., Takarada, T., Saito, I., 2008. Low tem- Process. 46, 630–636.
perature catalytic steam gasification of HyperCoal to produce H2 and synthesis Zoulalian, A., Bounaceur, R., Dufour, A., 2015. Kinetic modelling of char gasification
gas. Fuel 87 (4–5), 491–497. by accounting for the evolution of the reactive surface area. Chem. Eng. Sci. 138,
Silbermann, R., Gomez, A., Gates, I., Mahinpey, N., 2013. Kinetic studies of a novel 281–290.

Potrebbero piacerti anche