Sunteți pe pagina 1din 14

energies

Article
High Quality Syngas Production with Supercritical
Biomass Gasification Integrated with a Water–Gas
Shift Reactor
M. M. Sarafraz 1 , Mohammad Reza Safaei 2,3, * , M. Jafarian 1 , Marjan Goodarzi 4,5, * and
M. Arjomandi 1
1 School of Mechanical Engineering, University of Adelaide, Adelaide, Australia
2 Division of Computational Physics, Institute for Computational Science, Ton Duc Thang University,
Ho Chi Minh City, Vietnam
3 Faculty of Electrical and Electronics Engineering, Ton Duc Thang University, Ho Chi Minh City, Vietnam
4 Department of Mechanical Engineering, Lamar University, Beaumont, TX 77705, USA
5 Sustainable Management of Natural Resources and Environment Research Group, Faculty of Environment
and Labour Safety, Ton Duc Thang University, Ho Chi Minh City, Vietnam
* Correspondence: cfd_safaei@tdtu.edu.vn (M.R.S.); mgoodarzi@lamar.edu or
marjan.goodarzi@tdtu.edu.vn (M.G.)

Received: 30 May 2019; Accepted: 1 July 2019; Published: 5 July 2019 

Abstract: A thermodynamic assessment is conducted for a new configuration of a supercritical


water gasification plant with a water–gas shift reactor. The proposed configuration offers the
potential for the production of syngas at different H2 :CO ratios for various applications such as the
Fischer–Tropsch process or fuel cells, and it is a path for addressing the common challenges associated
with conventional gasification plants such as nitrogen dilution and ash separation. The proposed
concept consists of two reactors, R1 and R2 , where the carbon containing fuel is gasified (in reactor
R1 ) and in reactor R2 , the quality of the syngas (H2 :CO ratio) is substantially improved. Reactor R1 is
a supercritical water gasifier and reactor R2 is a water–gas shift reactor. The proposed concept was
modelled using the Gibbs minimization method with HSC chemistry software. Our results show
that the supercritical water to fuel ratio (SCW/C) is a key parameter for determining the quality of
syngas (molar ratio of H2 :CO) and the carbon conversion reaches 100%, when the SWC/C ratio ranges
between two and 2.5 at 500–1000 ◦ C.

Keywords: supercritical water gasification; water–gas shift reactor; biomass gasification; syngas
quality

1. Introduction
The energy crisis and environmental pollution due to the combustion of carbon-containing fuels
have resulted in more research being conducted on renewable energy and clean energy resources
as potential alternatives to fossil fuels [1–3]. Synthetic gas (syngas) is a mixture of hydrogen and
carbon monoxide, which is a promising replacement for fossil fuels. It is a cheap, clean burning
fuel, easy-to-transport gas, and has a wide range of applications such as some types of fuel cell
systems [4,5] and the Fischer–Tropsch process [4]. Therefore, it has received ever-increasing special
attention throughout the past decades. Gasification is the common pathway to produce syngas from
carbonaceous fuels. In conventional air-blown gasification systems, air is used for the partial oxidation
of fuel, which not only adds impurities (e.g., sulphur dioxide) to the final gaseous products but supplies
nitrogen to the syngas, which reduces the quality of the syngas product. One plausible solution for
this challenge is to use oxy-fuel gasifiers to avoid the appearance of impurities and nitrogen at the

Energies 2019, 12, 2591; doi:10.3390/en12132591 www.mdpi.com/journal/energies


Energies 2019, 12, 2591 2 of 14

outlet, which in turn adds to the cost and complexity of the process. Therefore, there is a need to seek
alternatives that produce high quality syngas, while addressing the aforementioned challenges [5].
Chemical looping gasification is an emerging technology for the production of syngas using a
solid oxygen carrier (OC). This technology addresses the nitrogen dilution but also has the potential to
reduce and/store greenhouse gases (GHG) such as CO2 . In this concept, two interconnected reactors,
the gasifier and the air reactor, are employed. In the gasifier, metal oxides are reduced, and fuel is
partially oxidized [6]. Syngas is the main product of the gasifier. Then, the reduced metal oxides
are transported to the air reactor where particles of oxygen are recovered [7]. Notwithstanding the
advantages of the solid oxygen carrier particles, there are some challenges associated with the use of the
solid oxygen carriers in chemical looping systems. These include agglomeration and sintering [6,8–16],
and also the need to separate the OC particles from any carry-over particles from the gasifier, as well as
manage the deposition of carbon and ash on the OC particles [17]. These challenges significantly reduce
the effectiveness of the oxygen carrier particles to transport oxygen between the reactors [16,18–22]
and hence decreases the efficiency of the process [6]. However, this emerging concept is in the early
stage of development and needs further investigation to understand its shortcomings. For example,
a material constraint due to the sintering, breakage, and corrosion of metals, is one of the challenges
associated with the use of molten metal in the chemical looping process.
One promising method to gasify a carbonaceous fuel is supercritical water (SCW) gasification.
This concept offers a wide range of benefits over the other concepts such as:

• Supercritical water has zero surface tension and most of the carbon-containing fuels are soluble in
it, and therefore diffusion and penetration of water in the carbon with insignificant mass transfer
resistance is plausible [23];
• The SCW gasification process is flexible with respect to the type of the carbonaceous fuel. For
example, different types of biomass, coal, or even municipal waste with various contents of
moisture and impurities are used as fuel sources [24];
• The required operating temperature for the gasification lies between 400 ◦ C and 1000 ◦ C depending
on the type of the fuel and quality of the syngas;
• The produced carbon dioxide easily separates from H2 using pressurized water;
• Some physical properties of water, such as density, ion product, dielectric constant, viscosity,
diffusivity, and solubility, change near or at its thermodynamic critical point (T = 374 ◦ C and P =
22.1 MPa). At the critical point, water behaves similar to a dense gas with a consequent removal of
any interphase mass transport processes. Organic compounds have high solubilities and complete
miscibility with supercritical water [25];
• The process is high pressure, which reduces the costs related to the storage of the gaseous products
such as post-compression operation.

These advantages provide plausible conditions for better gasification of carbonaceous fuel,
particularly biomass in supercritical water. For example, thermodynamic assessment and system
modelling for the gasification of biomass with supercritical water were conducted by Withag et al. [26].
The selected feedstock was a wet biomass comprised of 70% water by weight. It was shown that the
gasification is feasible without any further drying process. Thermochemical equilibrium analysis was
used for the modelling of the process and it was found that the composition of the product gases could
be tailored to the desired product composition by changing the process parameters such as the reactor
temperature, pressure, and the concentration of organic material in the feed. However, the pressure
of the reactor was 100–300 bar, which was technically challenging and expensive. In addition, the
proposed system produced 38% CO2, which was one main challenge of the supercritical gasification at
higher temperatures.
In another work, Guan et al. [27] investigated the gasification of algae nanochloropsis in
supercritical water and showed that with an increase in the temperature of the gasifier, more carbon
dioxide is produced. By reducing the operating temperature of the gasifier, the mole fraction of
Energies 2019, 12, 2591 3 of 14

CO2 slightly decreased, however, the quality of the produced synthetic gas (H2 :CO molar ratio)
was low. A similar trend was reported in the gasification of dry starch as biomass conducted by
Yakaboylu et al. [28]. Thus, to achieve high quality syngas, there is a need to separate the CO2 from
other gaseous products. In a study conducted by Guo et al. [29], they reported the controversial result
that operating temperature has a strong influence on the gasification of biomass with supercritical water
such that the gasification efficiency and H2 production at higher temperatures inverses the quantity
of the CO2 product [29]. Thus, further investigation on the role of temperature on the supercritical
gasification of biomass and mole fraction of gaseous products is required.
The diversity of components in gaseous products obtained by the gasification pathway has also
been a popular subject of research, because carbon-containing fuels such as biomass have different
compositions of cellulose, glucose, glycerol, lignin, and phenolic which may result in a substantial
change in the gasification reactions and consequently changes the configuration of the reactor [29–31].
For example, Guo et al. [32] performed an experimental investigation on the catalytic and noncatalytic
supercritical gasification of glycerol in a tubular quartz reactor for hydrogen production. They showed
that by using a catalyst, the main gaseous products are hydrogen with the mole fraction of 59%,
followed by CO2 with the mole fraction of 29.9%, CH4 with the mole fraction of ~6.5%, and CO with the
mole fraction of ~4.5%. Noticeably, in the absence of the nickel catalyst, the mole fraction of hydrogen
decreased to 50%, the mole fraction of CO2 decreased to 24.93%, and the mole fraction of CO decreased
to 21.13%. Therefore, depending on the type of feedstock, a specific type of the reactor or configuration
of the process is required [33–35].
For example, one potential configuration for supercritical gasification is a fluidized bed system.
In a study conducted by Lu et al. [36], the hydrodynamic behavior of a supercritical gasifier was
evaluated at temperatures ranging from 360 ◦ C to 420 ◦ C and pressure ranging from 23 MPa to 27 MPa.
They identified that a double symmetric feeding pipe with an angle of 45◦ provided uniform solid
distribution and a long residence time, which potentially also represented better chemical efficiency.
However, it was identified that depending on the composition of the feedstock, the configuration of
the system might need a major modification in order to produce high-quality syngas [33]. Hence,
in our research, the thermodynamic potential for a new configuration that would produce a high
quality syngas with supercritical water is investigated for graphite (pure carbon) as a surrogate for any
carbonaceous fuel and biomass. The presence of the water–gas shift reactor removes the barrier of
dependence of the configuration of supercritical gasification on the composition of feedstock. Thus,
the proposed gasification plant is combined with a water–gas shift (WGS) reactor to control the H2 :CO
ratio. The influence of different operating conditions, including the ratio of supercritical water to
feedstock, temperature, and the pressure of the reactors on the gaseous products is investigated. The
chemical performance of the proposed concept is investigated for three different biomass feedstocks.

2. Methodology
Figure 1 presents a schematic diagram of the operating conditions considered for this research.
To apply the thermochemical equilibrium and sensitivity analysis, different temperatures and pressures
were applied to reactors R1 and R2 . The minimum temperature and pressure to achieve the supercritical
water was 374 ◦ C and 25 bar, respectively, which were applied to reactor R1 . For storing the syngas,
it was plausible to pressurize the Reactor R2 . A sensitivity analysis on the quality of the syngas and the
temperature of reactor R2 was conducted to identify the optimum range of the temperature for the
reactor, which was between 500 ◦ C and 800 ◦ C. At this range of temperature, the mole fraction of CO2
and methane was also minimized, which increased the quality of the syngas. In this research, a syngas
quality (H2 :CO molar ratio) greater than 2 was the target value of the simulation, which potentially
has a wide range of applications in gas to liquid processes, transportation fuels, the Fischer–Tropsch
process, and fuel cells. Thus, the last stage was to remove any moisture content from the syngas using
a refrigerant cooler.
Energies 2019, 12, 2591 4 of 14
Energies 2019, 12, x FOR PEER REVIEW 4 of 15

Figure 1. Detailed operating conditions of the process for supercritical synthetic gas (syngas) production.
Figure 1. Detailed operating conditions of the process for supercritical synthetic gas (syngas)
Figure 2 presents a schematic diagram of the water supercritical gasification, consisting of two
production.
reactors namely the supercritical gasifier (R1 ) and the WGS reactor (R2 ). To efficiently use the released
Figure
heat from the2exothermic
presents a reactions
schematicindiagram of the
the gasifier water supercritical
(methanation and WGSgasification, consisting
reactions), both of two
water (stream
reactors
1), namely
and2019,
Energies fuel x FORthe
(stream
12, supercritical
2) were
PEER fed into gasifier
REVIEW (R1) andgasifier
the supercritical the WGS reactor
(reactor R1 ).(R2). To efficiently use
5 ofthe
15
released heat from the exothermic reactions in the gasifier (methanation and WGS reactions), both
water (stream 1), and fuel (stream 2) were fed into the supercritical gasifier (reactor R1).
To enhance the quality of syngas (H2:CO molar ratio), the gaseous products from R1 (stream 7)
were fed into a WGS reactor (reactor R2). A cold water (stream 4) was fed into the reactor R2
proceeding the WGS reaction, which is exothermic and absorbs the released heat. The outlet from R2
(stream 8) was fed into a heat exchanger to produce a high-pressure steam (stream 12) from the cold
water pumped into the heat exchanger (stream 3). The generated high-pressure steam was then fed
into a two-stage steam turbine to produce ~11% of the total input energy to be used as electrical
power for work in the plant. Then, the cold syngas from the heat exchanger (stream 10) was stored
for further use. Notably, to the best of our knowledge there is no industrial WGS reaction, or any
demonstration of it, which operate at the proposed conditions yet. However, the scope of this work
is to thermodynamically assess the potential of the proposed system for high quality syngas
production and an investigation on this subject is beyond the scope of the present investigation.

Figure 2. Schematic diagram of the supercritical gasification.


Figure 2. Schematic diagram of the supercritical gasification.

To enhance the quality of syngas (H2 :CO molar ratio), the gaseous products from R1 (stream 7)
To predict the potential reactions occurring in the reactors, the Gibbs minimization method [37–
were fed into a WGS reactor (reactor R2 ). A cold water (stream 4) was fed into the reactor R2 proceeding
39] was employed. To achieve this, the Aspen Plus RGibbs reactor and HSC chemistry were used to
the WGS reaction, which is exothermic and absorbs the released heat. The outlet from R (stream 8) was
estimate the Gibbs free energy of the potential reactions. To solve the model, 2 the following
fed into a heat exchanger to produce a high-pressure steam (stream 12) from the cold water pumped
assumptions were considered:
into the heat exchanger (stream 3). The generated high-pressure steam was then fed into a two-stage
1.
steamThe reactions
turbine reach to~11%
to produce the equilibrium;
of the total input energy to be used as electrical power for work in the
2. Heat
plant. lossthe
Then, is cold
negligible
syngasfrom
fromallthe
reactors, pipes, tanks,
heat exchanger and10)
(stream units;
was stored for further use. Notably,
3. Heat
to the bestand mass
of our transfer coefficients
knowledge there is noare plausible
industrial to maintain
WGS reaction,the
or conditions for highest
any demonstration of chemical
it, which
performance of the reactions;
4. Graphite in the present work is a surrogate for more realistic feedstock, which is used in the
Gibbs minimization simulation. Any impurities in the feedstock have a negligible influence on
the reactions and only carbon reacts in the supercritical gasifier. If there is an impurity, it is
completely separated in the form of ash from the supercritical gasifier due to the difference
Energies 2019, 12, 2591 5 of 14

operate at the proposed conditions yet. However, the scope of this work is to thermodynamically
assess the potential of the proposed system for high quality syngas production and an investigation on
this subject is beyond theFigure scope2.of the present
Schematic diagram investigation.
of the supercritical gasification.
To predict the potential Figure 2. Schematic
reactions occurring diagram of the
in the supercritical
reactors, the Gibbs gasification.
minimization method [37–39]
To predict To
was employed. theachieve Figure
potential 2. Schematic
reactions
this, the Aspen diagram
occurring Plus in of the supercritical
the
RGibbs reactors,
reactorthe andgasification.
GibbsHSCminimization
chemistry were methodused[37– to
To predict the potential Figure 2. Schematic
reactions occurring diagram in theof the supercritical
reactors, the Gibbs gasification.
minimization method [37–
39] wasthe
estimate employed.
Gibbs free Toenergy
achieve of this, the Aspen
the potential Plus RGibbs
reactions. To solve reactor the model, and HSC chemistryassumptions
the following were used to
To predict the potential Figure 2. Schematic
reactions occurring diagram of the
in the supercritical
reactors, the and gasification.
Gibbs minimization method [37–
39] was
estimate
were employed.
the Gibbs
considered: Tofreeachieve energy this,ofthethe AspenpotentialPlus RGibbs
reactions. reactor To solve HSC the chemistry
model, were
the used
following to
To predict the Figurereactions2. Schematic diagramin ofthe
the reactors,
supercritical gasification.
39] was
estimate
assumptions employed.
the were To
Gibbsconsidered:free energy of the potential reactions. To solve the model, the following[37–
potential
achieve this, the AspenoccurringPlus RGibbs reactor the
and Gibbs
HSC minimization
chemistry were method
used to
39]
1.estimate To
was predict
employed.
the were
The reactions Gibbs the
reach potential
To
free toachieve
energy reactions
the equilibrium;this,
of the theoccurring
Aspen Plus
potential inreactions.
the reactors,
RGibbs Tothe
reactor solveGibbs
andthe minimization
HSC chemistry
model, the method
were used
following [37–to
assumptions To predict considered:
the potential reactions occurring in the reactors,
1.39] wasreactions
The
estimate employed. reachTo freeachieve
tofromthe this, the
equilibrium; Aspen Plus RGibbs reactor and HSC chemistry were used[37–
the Gibbs minimization method to
2.assumptions
Heat lossthe
were Gibbs
considered:
is negligible energy of the
all reactors, potential
pipes, tanks,reactions.
and units;To solve the model, the following
1. 39] was
The
estimate employed.
reactions
the
2. Heat loss iswere Gibbs reach
negligible To to
free achieve
the
energy this,
equilibrium;
of the
the Aspen
potential
from all reactors, pipes, tanks, and units; Plus RGibbs
reactions. reactor
To and
solve HSC
the chemistry
model, the were used
following to
3.1. assumptions
Heat
The
estimate andthe
reactions mass reach
Gibbs
considered:
transfer to
free coefficients
the equilibrium;
energy of arethe plausible
potential to maintain
reactions. the To conditions
solve the formodel,
highestthe chemical
following
2. Heat
3.assumptions loss
Heat and mass is negligible
were transferconsidered: from all reactors, pipes, tanks, and units;
coefficients are plausible to maintain the conditions for highest chemical
performance of the reactions;
3. 1.
2. Heat Theloss
assumptions
Heat and
performance
reactions
is
massnegligible
were of
reach
considered:
transfer
the
to
from the allequilibrium;
coefficients
reactions; reactors, pipes, tanks,
are plausible and units;
to maintain the conditions for highest chemical
4.3. 1.2.Graphite
The
Heat Heat reactions
and in
loss the
mass is reach
present
negligible
transfer toworkthefrom equilibrium;
is
coefficients a surrogate
all reactors,
are for more
pipes,
plausible to realistic
tanks,
maintain and feedstock,
units;
the conditionswhich for is used
highest in the Gibbs
chemical
performance
4.2.1.Graphite
The loss inisthe
reactions of thereachreactions;
present work
tofrom
the is reactors,
a surrogate
equilibrium; fortanks,
more and realistic feedstock, which is used in the
3. Heat
minimization
Heat and negligible
simulation.
mass transfer Any all impurities
coefficients are pipes,
in the
plausible feedstock
to maintain units;
have thea negligible
conditions influence
for highest on the
chemical
4. 2.performance
Graphite
Gibbs
Heat in
minimization
loss
of the
the
is
reactions;
present
negligible work
simulation.from is
all a
Any surrogate
impurities
reactors, for
pipes, more
in the
tanks, realistic
feedstock
and units; feedstock,
have a which
negligible is used
influence in the
on
3. Heat
4. reactions and
performance
Graphite andin mass
only
the transfer
ofcarbon
the
present coefficients
reactsisinathe
reactions;
work are
surrogate plausible
supercritical for to
more maintain
gasifier. realisticIf there the conditions
is an impurity,
feedstock, which for highest
it is
is used chemical
completely
in the
Gibbs
the
3.separated minimization
reactions
Heat and and
mass onlysimulation.
transfer carbon Any
reacts impurities
in the in
supercriticalthe feedstock gasifier. haveIf a
there negligible
is an influence
impurity, it onis
4. performance
GibbsGraphite in the
minimization in of
form
the the of
present ashcoefficients
reactions;
simulation. fromAny
work the are plausible
is asupercritical
surrogate
impurities for
in
to
gasifier
themore maintain due
feedstock realistictothe the conditions
difference
feedstock,
have a negligible
for highest
between
which thechemical
is used
influence ashinisthe
on
the reactions
completely and
separated only
the in carbon
the formreacts in the supercritical gasifier. If there is an impurity, it
4. and
the
performance
Graphite
supercritical
Gibbs
reactions in theof
minimization
and present
water
only
reactions;
work
density;
simulation.
carbon is aof
reacts Any
ash fromfor
surrogate
in impurities
the
themore
supercritical
supercritical
in therealistic
feedstock
gasifier.
gasifier due
feedstock,
Ifhave
theredue a is whichto the
negligible
an is difference
impurity,used in
it isthe
influence on
completely
between
4. Gibbs
Graphite the separated
ash
in and
the presentin the
supercritical form
work of
water
is a ash from
density;
surrogate the
for supercritical
more realistic gasifier
feedstock, to
which the difference
is used in the
5. The the minimization
reactions
residence time and simulation.
only
is sufficient carbonfor the Any
reacts impurities
of reactions in thetothe in the
supercritical
reach feedstock
completion gasifier. have
so that a
If there negligible
is an
no unreacted influence
impurity,
carbon it is on
5. completely
between
The residence
Gibbs
the
the
reactions
separated
ash
minimization and
time
and only
insufficient
the
supercritical
is simulation.
carbon
form water
for
reactsAny
ash
thein from
density;
reactions
impurities
the
supercritical
to the
in
supercritical reach gasifier
completion
feedstock
gasifier. have
If there
due
so to
that
a negligible
is an
the
no difference
unreacted
influence
impurity, it ison
enters
5. between
The completely
the WGS
the ash
residence separated
reactor;
and
time in
supercritical
is sufficient the form
water
for of ash
thedensity; from
reactions the supercritical
to reach completion gasifier so thatdue to the difference
no impurity,
unreacted
carbon
the
completelyenters
reactions the WGS
and
separated only reactor;
carbon
in the form reacts in the supercritical gasifier. If there is an it is
6.5. No between
catalytic the
effect ash is and supercritical
considered infortheof waterash
modelling. from However,
density; the to supercritical
reachmetal gasifier
oxides and due some tonothe difference
composites
6. The carbon
No residence
enters
catalytic
completely
time
the
effect WGS
is
separated
is sufficient
reactor;
considered in the the reactions
modelling. However, completion
metal oxides so
and that
some unreacted
composites
6. 5.have
between
The
carbon
No beenenters
catalytic
thethe
residence
identified ash
effect
and
time
isas
WGS is in
suitable
reactor;
considered
the form
asupercritical
sufficient heat
in the
water
for
and
of the
ash
mass
modelling.
from the to
density;
reactions
transferHowever,
supercritical
medium reach metal [40–44],
gasifier so
completion
oxides which and can
duethat
some
to the
also no difference
unreacted
improve
composites
have been
between identified
the ash and as a suitable
supercritical heat
water and mass
density; transfer medium [40–44], which can also
6. 5. the
No The
havecarbon
improve
residence
supercritical
catalytic
been the
enters
effect
identified
time
the
supercritical
WGS
gasification is sufficient
is considered
as a reactor;
reactions
suitable
gasificationin the for[45–47];
heat
the reactions
modelling.
and
reactions mass However,
[45–47];
to reach
transfer metal completion
medium oxides and
[40–44],
so somethat no
which
unreacted
composites
can also
5. carbon
Thecatalytic
residence
enters thetime WGS isreactor;
sufficientinfor the reactionsHowever, to reach metal completion so that nocomposites
unreacted
7.7. 6.The
have No
improve
The been
process
process the identified
isis effect
isobar
supercritical
isobar
is
andasconsidered
and agasification
the
thesuitable
pressure
pressure
thethe
heat
of modelling.
and
reactions mass
reactors
ofmodelling.
the [45–47];
reactors transfer
isisthe thesame. medium oxides
same. [40–44], and some
which can also
carbon
6. improve
No catalytic enters effectthe isWGSconsideredreactor; in the However, metal oxides and some composites
7. 6.ThehaveNoprocess been
the
catalytic
identified
supercritical
is isobar
effect and as
theapressure
is considered
suitablereactions
gasification
in ofheat
the the and
reactors
modelling.
mass
[45–47]; is the transfer
However, same.metal medium [40–44], which can also
oxides and whichsome composites
Table
have
Table 1 1shows
improve been
shows thethethe potential
identified potential
supercritical as reactions
a suitable
reactions
gasification occurring
heat
occurring and
reactions inin the
massthe proposed
[45–47]; transfer
proposed concept.
medium
concept. [40–44], can also
7. The process is isobar and the pressure of the reactors is the same.
have
Table
improve1 showsbeen
the identified
the potential
supercritical as a suitable
reactions
gasification
7. The process is isobar and the pressure of the reactors is the same. heat
occurring
reactionsand in mass
the
[45–47]; transfer
proposed medium
concept. [40–44], which can also
Table 1
improve shows
7. The process is isobar the the potential
supercritical
Table
Table and1. reactions
Potential gasification
reactions
the pressure
1. Potential occurring
reactions reactions
of
of of in
supercritical
the the
reactors iswater
supercritical proposed
[45–47]; water the same. with concept.
carbon.
with carbon.
7. Table
The process 1 shows the
isTable
isobar potential
1.and thereactions
Potential pressure
reactions occurring
of
of the in thewater
reactors
supercritical proposed
is thewith same. concept.
carbon.
Table No. 1No. shows the Table potential Name
Name
1. Potential reactions
reactions occurring in the proposed
of supercritical waterReaction Reaction
withconcept.
carbon.
Table No. 111 shows theTable potential Name
Reforming
1. reactions
Potential
Reforming occurring
reactions CinC+the
of supercritical +HH proposed
2O(g)
2 O(g)water ↔with
Reaction HHconcept.
(g)
22(g) ++CO(g)
carbon. CO(g)
No. 122 Table
Partial 1. Name
Reforming
Potential
oxidation
Partial oxidation of graphite of reactions
graphite of C
supercritical
3C ++
3C + 2H2O(g) H2H 2 O(g)
2 water
O(g) ↔
Reaction
with H
CH (g)
carbon.
CH4(g) +2CO(g)
24 (g) ++ CO(g)
2CO(g)
1233 No.Complete Complete
Partial Table Reforming
oxidation
oxidation
Nameof reactions
1. Potential
oxidation of of graphite
graphite
graphite
of 3C
C C +
+ +
+
supercritical
C 2H
2H H
2H 2O(g) ↔ 2H
22O(g)O(g)
O(g)
2 water ↔ Reaction
H
with
2H
CH 22(g)
(g) (g)++++CO(g)
carbon.
24(g)
CO 2 (g)
2CO(g)
CO 2(g)
No.
2344 1 Complete
Name
Reforming
Methanation C +
C +
H O(g)
22H ↔ (g) ↔
Reaction CHH (g)
(g) + CO(g)
Partial oxidation of graphite 3C + 2H O(g) ↔CH (g) + 2CO(g)
2 24
oxidation of graphite C + 2H 22O(g) 2H 4(g) + CO 2(g)
↔↔
Methanation C + 2H (g) 2CH (g)
No.
3455 1 2 Complete Water-gas
Partial
Name
Reforming
shift of
oxidation reaction
of graphite CCO(g) +3CC +++ H
22H
H2O(g)
2
O(g)
22O(g) ↔
Reaction

HH CH2(g)
4
(g)4++(g)
+COCO(g)
CO (g)
+222CO(g)
Water-gas oxidation
Methanationshift graphite
reaction CO(g) 2H C + O(g)
+HH 2H
2O(g) 2(g)
+ ↔↔ ↔2H 22
H (g)
CH 2(g) 4(g)+ CO (g)
2(g)
456
6 2 31 Complete Boudouard
PartialMethanation Reforming
oxidation
oxidation reaction
of graphite
of graphiteCO(g) 3CC+C C CO
++2H +2H
+H222H
(g)
2O(g)
2 O(g)
2(g) ↔
2O(g)
C ↔CH 2H
CH
H
2CO(g) (g)
4(g)
2
24(g)
(g)
+ CO(g)
++2CO(g)
CO 2(g)
2 Water-gas
Boudouard
Partial shift
oxidation reaction
reaction of graphite 3C CO
+ 2H 2O(g)
(g) 2O(g)↔ 2H
+ C ↔ H

2(g)
2CO(g)
CH +
4(g)
CO+ 2(g)
2CO(g)
+ C2↔
3 Complete oxidation of graphite C + 2H O(g) (g) + CO (g)
56 4 Methanation + H2C(g) + 2H ↔H
(g) CH 4(g)2(g)
2 2 2
Water-gas shiftreaction
reaction CO(g)CO 2O(g) 2(g) + CO
4 53chemical
Boudouard
Complete oxidation of graphite C following
+C 2H 2O(g) ↔ ↔ 2CO(g)
2H 2(g) ++CO CO2(g)
2(g) + C ↔ 2CO(g) defined:
To assess 6the Water-gasMethanation
performance
Boudouard shift
reactionof reaction
the reactors, CO(g)
the COfollowing+ + H2H 2 2(g)
O(g) parameter CH
H 2 4(g)
(g)
is 2(g)
To assess the 56 4 chemical Water-gasperformance
Methanation
Boudouard shiftreactionof
reactionthe reactors, the
CO(g) +CO C +
H22O(g) 2H
(g) 2
+(g)
C ↔
↔↔H22CO(g)
parameter CH is
(g)
(g) + CO2(g)
4 defined:
To assess the5 chemical
𝑆𝐶𝑊 6 𝑛
performance
Water-gas shift of
SCW the reactors,
reaction nsteamCO(g) the following +2H (g)2O(g) ↔ 2CO(g)
parameter
↔ H2(g)is+defined: CO2(g)
To assess the chemicalBoudouard performance reaction
of the = reactors, the following CO + Cparameter is defined: (1)
𝑆𝐶𝑊
To 𝐶
𝑛
assess𝑛the chemical performance of the reactors,
6 Boudouard C
reaction n CO 2 (g) +
f eedstock the following parameter is defined: C ↔ 2CO(g) (1)
𝑆𝐶𝑊 (1)
𝐶
To assess 𝑛𝑛 chemical performance of the reactors, the following parameter is defined: (1)
the
where, n steam 𝐶 𝑆𝐶𝑊
is the𝑛 moles 𝑛 of supercritical steam required for the gasification and n feedstock is the moles
To assess the chemical performance of the reactors, the following parameter is defined: (1)
𝑆𝐶𝑊
of feedstock fed𝐶into the 𝑛
𝑛 reactor. SCW and C stand for supercritical water and carbon, respectively. The(1)
𝑆𝐶𝑊
𝐶 𝑛
carbon conversion is𝑛 also defined with the following equation: (1)
𝐶 𝑛
n f eedstock,initial − n f eedstock,remaining
x= × 100 (2)
n f eedstock,initial

where, n feedstock, initial is the moles of feedstock introduced to the reactor and n feedstock, remained is the
unreacted moles of carbon or feedstock. To perform the sensitivity analysis and to compare the results
to a reference case, a reference condition is defined in Table 2.
To validate the results of the modelling, a comparison was made between the results obtained
with the developed model and those reported in the literature. To make this comparison, the model
𝑛 ,

where, n feedstock, initial is the moles of feedstock introduced to the reactor and n feedstock, remained is the
unreacted moles of carbon or feedstock. To perform the sensitivity analysis and to compare the
results to a reference case, a reference condition is defined in Table 2.
Energies 2019, 12, 2591 6 of 14
Table 2. Reference conditions used in the sensitivity analysis.

was validated using Parameter


the same operating T R1conditions
(°C) T R2given
(°C) in Pressure (bar) The
the literature. SCW/C results of comparison
showed that theRange (min–max)
estimated moles of650–1000
production500–800
of hydrogen25–60 was similar 0.01–3
to those reported in the
literature [48–50]. Reference
As showncase in Figure 650 600of production
3a, the moles 25 of hydrogen 1 estimated with the
model is within the deviation of ±10% of those of reported in the literature [51,52]. Likewise, the syngas
qualityTo(H
validate the results of the modelling, a comparison was made between the results obtained
2 :CO molar ratio), as presented in Figure 3b, was in good agreement with the literature with
awith the developed model and those reported in the literature. To make this comparison, the model
deviation of ±15%.
was validated using the same operating conditions given in the literature. The results of comparison
showed that the estimatedTable moles of production
2. Reference conditionsof hydrogen
used was similar
in the sensitivity to those reported in the
analysis.
literature [48–50]. As shown in Figure 3a, the moles of production of hydrogen estimated with the
T R1 (◦ C) R2 (◦ C)
model is within Parameter
the deviation of ±10% of thoseTof reported Pressure (bar)
in the literature SCW/C
[51,52]. Likewise, the
syngas quality Range (min–max)
(H2:CO 650–1000
molar ratio), as presented 500–800
in Figure 3b, 25–60 0.01–3
was in good agreement with the
literature with Reference case
a deviation of ±15%. 650 600 25 1

(a) (b)
the model
Figure 3. Validation of the model using
using the
the data
datareported
reportedininthe
theliterature,
literature,(a)
(a)comparison
comparisonbetween
between
estimated moles of production of hydrogen and data reported in previous works [48–50], (b)
estimated moles of production of hydrogen and data reported in previous works [48–50], (b) comparison
between
comparisonthe estimated
between thesyngas qualitysyngas
estimated (H2 :CO molar(H
quality ratio)
2:COand those
molar reported
ratio) in the reported
and those literaturein
[51,52].
the
literature [51,52].
To assess the effect of biomass composition on the composition of the gas production, three
different biomasses
To assess were selected
the effect of biomassfrom the literature.
composition on The proximate analysis
the composition of the and
gas ultimate analysis
production, threeof
the biomasses are listed in Table 3. It is worth noting that most of the biomass have a very
different biomasses were selected from the literature. The proximate analysis and ultimate analysis low sulphur
content (<1%) andare
of the biomasses a high
listedoxygen content
in Table 3. It is(>35%). The content
worth noting of sulphur
that most of the considerably
biomass haveinfluences
a very lowthe
quality
sulphurofcontent
the syngas.
(<1%)The andlower the oxygen
a high sulphur content
content,(>35%).
the higher
Thethe qualityofofsulphur
content the syngas [53–55].
considerably
influences the quality of the syngas. The lower the sulphur content, the higher the quality of the
syngas [53–55]. Table 3. Analysis of the agro biomass feedstock used in this research [56].
Feedstock Proximate Analysis Ultimate Analysis
Table 3. Analysis of the agro biomass feedstock used in this research [56].
Moisture Volatile Fixed
Biomass Ash C H N S O
Feedstock Content Proximate analysisCarbon
Matter Ultimate analysis
Biomass
Feedstock Moisture
1 55.31 Ash
60.32 Volatile 9.36
30.32 Fixed
22.3 3.3 C 2.1 H 0.5N S
11.3 O
Feedstock 2 47.98 40.24 34.55 25.21 50.2 3.8 2.7 0.5 2.4
Feedstock 3 41.39 52.67 32.19 15.14 37.8 3.1 2.3 0.42 3.6

3. Results and Discussion

3.1. Gibbs Free Energy and Enthalpy Assessment


Figure 4a presents the dependence of the Gibbs free energy on the operating temperature of
the gasifier and the WGS reactor for the different reactions listed in Table 1. As shown in Figure 3a,
at T > 650 ◦ C, the Gibbs free energy for reactions 4 (methanation) and 5 (partial oxidation of graphite)
is positive, meaning that these reactions are unlikely to occur at this temperature range. However,
3.1. Gibbs Free Energy and Enthalpy Assessment
Figure 4a presents the dependence of the Gibbs free energy on the operating temperature of the
gasifier and the WGS reactor for the different reactions listed in Table 1. As shown in Figure 3a, at T
Energies 2019, 12, 2591 7 of 14
> 650 °C, the Gibbs free energy for reactions 4 (methanation) and 5 (partial oxidation of graphite) is
positive, meaning that these reactions are unlikely to occur at this temperature range. However, the
restrest
the of the reactions
of the reactions occur
occur within
within thethe
gasifier
gasifierandandthe
theWGS
WGSreactors.
reactors.Therefore,
Therefore,for for reactions
reactions to be
spontaneous and feasible in the gasifier, gasifier, the
the minimum
minimum temperature
temperature should should bebe at
at least 650 ◦°C,
least 650 since
C, since
temperature, the
over this temperature, the Gibbs
Gibbs free
free energy
energy for
for reactions
reactions 1,
1, 2,
2, 33 and
and 66 is
is negative.
negative. Figure 4b presents
the dependence of of enthalpy
enthalpyof ofreactions
reactionson onthe
thetemperature
temperature ofofthethe
gasifier and
gasifier thethe
and WGS WGSreactors for
reactors
the the
for different reactions
different reactions presented
presentedin Table 1 and
in Table shows
1 and thatthat
shows the the
enthalpy
enthalpyof the reactions
of the doesdoes
reactions not
significantly
not change
significantly withwith
change temperature
temperatureand remains approximately
and remains approximately constant. For example,
constant. for eachfor
For example, of
the reactions
each 1, 2, 3, and
of the reactions 1, 2,6,3,the
andenthalpy of reaction
6, the enthalpy remains remains
of reaction constantconstant
at a temperature range between
at a temperature range
650 °C and
between 650 ◦1000
C and °C.1000 ◦ C.
It is worth noting
It is worth that that
noting the total enthalpy
the total enthalpy of of
thethe
reaction
reactionfor forthe
the gasifier
gasifier is
endothermic and
endothermic and itit is
is exothermic
exothermic for for the
the WGS
WGS reactor.
reactor.

(a)

(b)
Figure 4. Dependence on temperature
4. Dependence temperature of the Gibbs free energy and enthalpy of the reaction
reaction with
with
temperature.
temperature. (a) Variation of the change in the Gibbs free energy of reactions with temperature the
(a) Variation of the change in the Gibbs free energy of reactions with temperature for for
reactions given in Table 1 and (b) variation of the enthalpy of the reaction with temperature for the
reactions given in Table 1.

3.2. Thermochemical Equilibrium Assessment for Reactor R1


Figure 5 presents the dependence of moles of gaseous products on the temperature of reactor
R1 for the supercritical water to fuel ratio (SCW/C) = 1 at P = 25 bar. It is worth noting that the
supercritical water gasification occurs at operating pressures larger than 25 bar. For P < 25 bar, the
supercritical gasification does not occur, and the chemical conversion of carbon is very low. For an
atmospheric gasification, more steam is required not only to enrich the hydrogen content but also to
provide sufficient mixing in order to drive the reaction towards completion. We observe that with an
increase in temperature the moles of production of H2 and CO increase. For example, at 700 ◦ C, the
mole fractions of H2 and CO are 0.86 kmol and 0.66 kmol, respectively. However, at T > 900 ◦ C, the
mole fractions o H2 and CO are 0.99 kmol and 0.97 kmol, respectively. In addition, with an increase in
temperature, the moles of production for CO2 and CH4 approaches zero. Importantly, this behaviour
is only seen for an SCW/C ratio of one and smaller. For example, for the SCW/C = 2, as presented in
atmospheric gasification,
provide sufficient mixing more steamtoisdrive
in order required not onlytowards
the reaction to enrich the hydrogen
completion. We content
observebut thatalso
withtoan
provide
increase sufficient mixing inthe
in temperature order
moles to drive the reaction
of production of H towards
2 and CO completion. Weexample,
increase. For observe that °C,anthe
with
at 700
mole fractions of H2 and CO are 0.86 kmol and 0.66 kmol, respectively. However, at T > 900 °C,
increase in temperature the moles of production of H 2 and CO increase. For example, at 700 °C, thethe
mole
mole fractions
fractions of oHH 2 and
2 andCO COareare0.86
0.99kmolkmoland and0.660.97kmol,
kmol,respectively.
respectively.However,
In addition, at Twith
> 900an°C, the
increase
mole
in fractions o H2the
temperature, and moles
CO areof0.99 kmol and for
production 0.97CO kmol,
2 and respectively.
CH In addition,
4 approaches zero. with an increase
Importantly,
Energies 2019, 12, 2591 8 this
of 14
in behaviour
temperature, the moles of production for CO and CH approaches
is only seen for an SCW/C ratio of one and smaller. For example, for the SCW/C = 2, as
2 4 zero. Importantly, this
behaviour
presented is in
only seen6,for
Figure at 700 °C, the moles
an SCW/C ratio ofofone and smaller.
production of H2For andexample,
CO are 1.36 for the
kmol SCW/C
and 0.54= 2,kmol,
as
◦ C, the moles of production of H and CO are 1.36 kmol and 0.54 kmol, respectively.
Figure
presented 6,
in at 700
Figure 6, at 700 °C, the moles of production of H and
respectively. In this case, the moles of production of CO2 is 0.38 kmol, which slightly decreases with
2 2 CO are 1.36 kmol and 0.54 kmol,
In increase
an this case,
respectively. inthe
In this moles
temperature. of production
case, the moles
The moles ofofCO
of production 2 is 0.38
production of CO kmol,
2 is
of CO which
0.38
2 atkmol, slightly
1000 which
°C decreases
slightly
is 0.25 with
kmol.decreases an increase
Importantly, withan
in temperature. The moles of moles
production oftheCOproduction ◦
anincrease
increase in temperature
in temperature. The
slightly of production
decreases 2 atof1000
CO2 C atofis
10000.25°Ckmol.
hydrogen Importantly,
is 0.25
and kmol. an
theincrease
Importantly,
increases molesan of in
temperature
increase in slightly
temperature decreases
slightly the production
decreases the of hydrogen
production of
production of CO, meaning that the quality of syngas (H2:CO molar ratio) slightly decreases. and
hydrogenincreases
and the moles
increases of
theproduction
moles of of
CO, meaning
production
Therefore, of that
theCO, the
chemical quality
meaning of syngas
that
performance the of (H
the2 :CO
quality ofmolar
reactor R1 ratio)
syngas (Hslightly
strongly 2:CO decreases.
molar
depends theTherefore,
ratio)
on slightly
molar the
ofchemical
decreases.
ratio SCW/C
performance
Therefore,
and ofdependent
the chemical
is slightly the reactor R1the
performance
on strongly depends
of the
temperature reactorofon R1the
the molarRdepends
strongly
reactor ratio
1. of SCW/C
on theandmolaris slightly
ratio of dependent
SCW/C
andonisthe temperature
slightly dependent of the
on reactor R1 .
the temperature of the reactor R1.

Figure 5. Variation of the mole fraction of the productions with temperature of the reactor R1 at
Figure 5. Variation
reference
Figure of given
5.conditions
Variation the mole
of the fraction
inmole
table 2. of of
fraction thethe
productions with
productions temperature
with of of
temperature thethe
reactor R1 R
reactor at1 at
reference conditions given in table 2.
reference conditions given in Table 2.

Figure 6. Variation of the mole fraction of the products with temperature at the reference condition in
reactor at the supercritical water to fuel ratio (SCW/C) = 2.

Figure 7 presents the dependence of the molar ratio of H2 :CO (referred to as syngas quality) on
the temperature of reactor R1 at the reference conditions. As shown in Figure 7, with an increase in the
molar ratio of SCW/C, the quality of syngas (H2 :CO molar ratio) increases, however, to achieve a target
value of 2.1, a higher operating temperature is required. For example, at a molar ratio of SCW/C = 2,
at T = 735 ◦ C, the H2 :CO molar ratio reached the target value, while for SCW/C = 3, the target value for
the H2 :CO molar ratio was satisfied at T = 988 ◦ C. Moreover, with an increase in temperature, moles of
production for CO2 decreases, hence two different regimes are seen in Figure 7. Importantly, as shown
in Figure 4a, the Gibbs free energy is negative for T > 650 ◦ C, and therefore it is likely that the reactions
proceed spontaneously in this range. Therefore, these two constraints limit the operating temperature
of reactor R1 between 650 ◦ C and 900 ◦ C. It is worth noting that at a large ratio of SCW/C, the energy
requirement of the system is intensified as more steam is demanded for the supercritical reactor.
target value for the H2:CO molar ratio was satisfied at T = 988 °C. Moreover, with an increase in
temperature, moles of production for CO2 decreases, hence two different regimes are seen in Figure
7. Importantly, as shown in Figure 4a, the Gibbs free energy is negative for T > 650 °C, and therefore
it is likely that the reactions proceed spontaneously in this range. Therefore, these two constraints
limit the operating temperature of reactor R1 between 650 °C and 900 °C. It is worth noting that at a
Energies 2019, 12, 2591 9 of 14
large ratio of SCW/C, the energy requirement of the system is intensified as more steam is demanded
for the supercritical reactor.

Figure7.7.Variation
Figure Variation ofof
the syngas
the syngas quality (H(H
quality 2 :CO
2:COmolar
molarratio) with
ratio) temperature,
with temperature, forfor
different values
different of
values
the supercritical water to fuel ratio (SCW/C) in reactor
of the supercritical water to fuel ratio (SCW/C) in reactor R 1 . The target (shown in the figure) is to
R1. The target (shown in the figure) is toobtain
the syngas
obtain thequality
syngas (H 2 : CO(H
quality molar
2: CO ratio)
molar value
ratio) of ~2.1,
value of suitable for the
~2.1, suitable forFischer–Tropsch
the Fischer–Tropschprocess and
process
some
and fuel
somecell applications.
fuel cell applications.

Figure
Figure8 8presents
presentsthe
thedependence
dependenceof offuel
fuelconversion
conversionon on temperature
temperature for for different molar ratios
different molar ratios of
of
SCW/C. AsAsshown, with anan
increase in in
SCW/C, carbon conversion increases. ForForexample, at 700 ◦ C,°C,
for
SCW/C. shown, with increase SCW/C, carbon conversion increases. example, at 700
SCW/C = 0.01, the fuel conversion is only 0.8%, while for SCW/C = 1, it is 69.8%. Thus,
for SCW/C = 0.01, the fuel conversion is only 0.8%, while for SCW/C = 1, it is 69.8%. Thus, the ratio of the ratio of SCW/C
strongly
SCW/Cinfluences the fuel conversion.
strongly influences Likewise, with
the fuel conversion. an increase
Likewise, with an inincrease
the temperature of the reactor,
in the temperature the
of the
fuel conversion
reactor, the fuelincreases, because
conversion with an
increases, increase
because in an
with theincrease
value ofinSCW/C,
the valuethe of
content
SCW/C, of H 2 Ocontent
the increases,
of
which drives Equations (1)–(4) including gasification, partial oxidation of carbon,
H2O increases, which drives Equations (1)–(4) including gasification, partial oxidation of carbon, methanation, and
complete oxidation
methanation, of carbon.oxidation
and complete These reactions
of carbon.cause
Thesemore carbon
reactions to bemore
cause consumed
carbonand to bethe value of
consumed
the example, for SCW/C =
and the value of the carbon conversion is promoted. For example, for SCW/C = 0.5 and SCW/C =C1,the
carbon conversion is promoted. For 0.5 and SCW/C = 1, at T = 800 ◦
at
fuel 800 °C theisfuel
T =conversion 45%conversion
and 90%, respectively,
is 45% and however, when the however,
90%, respectively, temperature whenis increased to 1000 ◦is
the temperature C,
the
increased to 1000 °C,
fuel conversion reaches 99.1%
the fuel and 100%,reaches
conversion respectively.
99.1% Therefore,
and 100%,itrespectively.
is estimated Therefore,
that graphite is
it is
completely
Energies 2019, 12,converted
estimated that PEERinto
x FOR graphite syngas
is completely
REVIEW > 1 if the required
at SCW/Cconverted temperature
into syngas at SCW/C is maintained.
> 1 if the 10required
of 15
temperature is maintained.

Figure
Figure8.8.Variation
Variationofofthe
thecarbon
carbon(fuel) conversion
(fuel) extent
conversion with
extent temperature,
with forfor
temperature, various values
various of of
values
SCW/C
SCW/Cratio
ratioininreactor
reactorRR
1.
1.

3.3.Thermochemical
3.3. ThermochemicalEquilibrium
Equilibrium Assessment
Assessment forfor Reactor
Reactor R2R2
Figure9 9presents
Figure presentsthe
thedependence
dependenceofofmole
mole fraction
fraction of of gaseous
gaseous products
products ononthethe temperature
temperature of of
reactor R for different components at P = 25 bar. As shown, with an increase in temperature,
reactor R22for different components at P = 25 bar. As shown, with an increase in temperature, the the mole
fraction
mole of Hof
fraction 2 and CO CO
H2 and increases. For For
increases. example, at Tat=T800
example,

= 800C,°C,
thethe
mole fractions
mole fractionsof of
H2Hand CO
2 and COare
are0.49
0.49 and 0.2, respectively, while at T = 1000 °C, they are 0.61 and 0.31, respectively. Moreover, an
increase in temperature decreases the mole fractions of CO2 and CH4, respectively. For example, at
700 °C, the mole fractions of CO2 and CH4 are 0.2 and 0.18, respectively, while at 1000 °C, they are
0.04 and 0.01, respectively. Therefore, we conclude that at higher temperatures, the WGS reactor
shows a better chemical performance. Notably, with an increase in the temperature of the water–gas
3.3. Thermochemical Equilibrium Assessment for Reactor R2
Figure 9 presents the dependence of mole fraction of gaseous products on the temperature of
reactor R2 for different components at P = 25 bar. As shown, with an increase in temperature, the
Energies 2019, 12, 2591 10 of 14
mole fraction of H2 and CO increases. For example, at T = 800 °C, the mole fractions of H2 and CO are
0.49 and 0.2, respectively, while at T = 1000 °C, they are 0.61 and 0.31, respectively. Moreover, an
increase
and 0.2,inrespectively,
temperature decreases
while at T = the
1000mole fractions
◦ C, they of CO
are 0.61 and2 and
0.31,CH 4, respectively.
respectively. For example,
Moreover, an increaseat
in °C,
700 the mole fractions
temperature decreasesofthe CO 2 and
mole CH4 areof0.2CO
fractions and 0.18, respectively, while at 1000 °C, they are
2 and CH4 , respectively. For example, at 700 C,

0.04
the and
mole0.01, respectively.
fractions of CO2 andTherefore,
CH4 arewe 0.2conclude
and 0.18, that at higherwhile
respectively, temperatures,
at 1000 ◦ C,thetheyWGS reactor
are 0.04 and
shows a better chemical performance. Notably, with an increase in the temperature
0.01, respectively. Therefore, we conclude that at higher temperatures, the WGS reactor shows a better of the water–gas
shift reactor,
chemical the production
performance. of CO
Notably, 2 decreases
with an increasesince the
in the water–gas of
temperature shift
the reactor
water–gas is an
shiftexothermic
reactor, the
reactor and an increase in the temperature of the reactor decreases the chemical
production of CO2 decreases since the water–gas shift reactor is an exothermic reactor and an conversion extent of
increase
the
in reaction. Moreover,
the temperature an reactor
of the increasedecreases
in the temperature
the chemical ofconversion
the reactor extent
causesofthethereaction
reaction. to Moreover,
proceed
inanthe reverse direction, which is endothermic. Normally, in the demonstration
increase in the temperature of the reactor causes the reaction to proceed in the reverse direction,cases for the WGS
reaction, the temperature is low (e.g., 200 °C) where the production of CO
which is endothermic. Normally, in the demonstration cases for the WGS reaction, the temperature
2 is maximized while theis
production of CO
low (e.g., 200 ◦ C) is suppressed,
where however,
the production increasing
of CO the temperature results in the promotion of the
2 is maximized while the production of CO is suppressed,
production of CO .
however, increasing the temperature results in the promotion of the production of CO2 .
2

Figure
Figure9.9.Variation
Variationofofthe
themole
molefraction
fractionofof
the product
the productwith
withtemperature
temperatureatatthe reference
the referencecondition
conditionfor
for
Energiesreactor
2019, RR
reactor
12, 2. 2
x . PEER REVIEW
FOR 11 of 15

Figure
Figure 1010presents
presents the
the dependence
dependence ofofmolemolefraction
fractionof gaseous products
of gaseous in syngas
products on theon
in syngas pressure
the
of reactor
pressure R2 at T R
of reactor =2800 ◦
at T C. According
= 800 to the LetoChatelier’s
°C. According principle,principle,
the Le Chatelier's with an increase
with an in pressure
increase in of
pressure of the reactor, the reactions proceed towards the production of more gaseous products tothe
the reactor, the reactions proceed towards the production of more gaseous products to reconcile
pressurethe
reconcile effect. Therefore,
pressure effect.increasing
Therefore,theincreasing
pressure hastheno influence
pressure onno
has theinfluence
chemical onperformance
the chemicalof the
WGS reactor [57]. However, pressurizing reactor
performance of the WGS reactor [57]. However, pressurizing R 2 can maintain the pressure required for the
reactor R2 can maintain the pressure syngas
or hydrogen
required for thestorage,
syngaswhich is advantageous.
or hydrogen Therefore,
storage, which there is no need
is advantageous. to add athere
Therefore, compressing
is no need unit
to to
thea plant,
add which reduces
compressing theplant,
unit to the cost significantly.
which reduces the cost significantly.

Figure 10.10.
Figure Variation of of
Variation thethe
mole fraction
mole of the
fraction products
of the with
products pressure
with at the
pressure reference
at the condition
reference in in
condition
reactor R2.R2 .
reactor

3.4. Biomass Composition


Figure 11a–c presents the dependence of mole fractions of gaseous products on different
carbon-containing fuels.
Figure
Energies 10.12,
2019, Variation
2591 of the mole fraction of the products with pressure at the reference condition in 11 of 14
reactor R2.

3.4.3.4. Biomass
Biomass Composition
Composition
Figure
Figure 1111a–c presents
a–c presents thethe dependence
dependence of of mole
mole fractions
fractions of of gaseous
gaseous products
products onon different
different
carbon-containing fuels.
carbon-containing fuels.

(a) (b)

(c)
Figure
Figure11.11. Mole
Mole percentage
percentage of of components
components in gaseous
in gaseous products
products forfor different
different molar
molar ratios
ratios of of SCW/C
SCW/C
and
and forfor different
different biomass
biomass feedstock
feedstock given
given in Table
in Table 2 and
2 and at at °C◦(a)
900900 C (a) feedstock
feedstock #1,#1,
(b)(b) feedstock
feedstock #2,#2,
and
and (c)(c) feedstock
feedstock #3.#3.

As shown in Figure 11a, feedstock#1, has the lowest amount of carbon content and the highest
amount of oxygen as compared with the other feedstocks. Therefore, more oxygen is available to
produce carbon dioxide and a lower syngas quality (H2 :CO molar ratio) is obtained as compared with
other feedstocks. For example, at SCW/C = 1, the syngas quality (H2 :CO molar ratio) is 1.02, 1.12, and
1.18, for feedstock 1, 2, and 3, respectively. It is worth noting that the low content of sulphur promotes
the quality of the syngas. In addition, the lower the content of nitrogen, the higher the syngas quality
(H2 :CO molar ratio). Feedstock#2, as presented in Figure 11b, has the lowest amount of oxygen content,
and therefore the mole fraction of CO2 is the lowest, while the quality of syngas (H2 :CO molar ratio)
of 2.31 is obtained at SCW/C = 2. Similarly, with an increase in the molar ratio of SCW/C, the mole
fraction of H2 increases for all the biomass feedstocks, while it decreases for CO2 and CO. For example,
for feedstock#1, with an increase in the SCW/C ratio from 1 to 2, the mole percentage of hydrogen
increases from 35.4% to 45.2%, while the mole percentage of CO and CO2 decreases from 34.5% to
29.9% and 29.03% to 23.85%, respectively. This is attributed to the increase in the content of hydrogen
and the suppression of reaction 3 due to the enrichment of hydrogen in the gas product. Since the
obtained results are based on the equilibrium Gibbs minimization model, a series of experiments is
required to validate and demonstrate the reactions and accurately justify the behaviour of the system
based on the kinetic parameters.
Energies 2019, 12, 2591 12 of 14

4. Conclusions
A thermodynamic assessment was conducted on a supercritical water gasification with a water–gas
shift reactor to assess the potential chemical performance of the proposed concept. The following
conclusions were made:

1. The addition of a water–gas shift reactor adds more control on the ratio of H2 :CO and minimizes
the production of CH4 and CO2 . For the proposed system, the quality of syngas (H2 :CO molar
ratio) reaches 2.1 at P = 25 bar and 850 ◦ C and 900 ◦ C for reactors R1 and R2 , respectively.
2. Pressure was found to have no influence on the chemical performance of the water–gas shift
reactor. However, pressurizing reactor R2 provides the pressure required for the storage of gas
and reduces the cost of post-compression of products.
3. The proposed system produces the syngas with different H2 :CO ratios depending on the SCW/C
and the temperature. For a given specific biomass, we found that syngas with the quality of 2.3
was produced at 900 ◦ C and SCW/C = 2.

Author Contributions: Conceptualization, M.M.S.; methodology, M.M.S., M.G., M.J., and M.A.; software, M.M.S.
and M.J.; validation, M.M.S. and M.J.; formal analysis, M.M.S., M.J., and M.R.S.; investigation, M.M.S., M.J., M.R.S.,
M.G., and M.A.; writing—original draft preparation, M.M.S., M.R.S., M.J., and M.G.; writing—review and editing,
M.M.S., M.J., M.R.S., M.G., and M.A.; supervision, M.A.
Funding: This research received no external funding.
Acknowledgments: The authors acknowledge the school of Mechanical Engineering, University of Adelaide for
the scientific support.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Turner, J.A. A realizable renewable energy future. Science 1999, 285, 687–689. [CrossRef] [PubMed]
2. Ogden, J.M.; Williams, R.H. Solar Hydrogen: Moving Beyond Fossil Fuels; World Resources Inst: Washington,
DC, USA, 1989.
3. Klass, D.L. Biomass for Renewable Energy, Fuels, and Chemicals; Academic Press: Cambridge, MA, USA, 1998.
4. Dry, M.E. The fischer–tropsch process: 1950–2000. Catal. Today 2002, 71, 227–241. [CrossRef]
5. Gemmen, R.; Trembly, J. On the mechanisms and behavior of coal syngas transport and reaction within the
anode of a solid oxide fuel cell. J. Power Sources 2006, 161, 1084–1095. [CrossRef]
6. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gayan, P.; Luis, F. Progress in chemical-looping combustion and
reforming technologies. Prog. Energy Combust. Sci. 2012, 38, 215–282. [CrossRef]
7. Jafarian, M.; Arjomandi, M.; Nathan, G.J. Thermodynamic potential of high temperature chemical looping
combustion with molten iron oxide as the oxygen carrier. Chem. Eng. Res. Des. 2017, 120, 69–81. [CrossRef]
8. Jafarian, M.; Arjomandi, M.; Nathan, G.J. A hybrid solar and chemical looping combustion system for solar
thermal energy storage. Appl. Energy 2013, 103, 671–678. [CrossRef]
9. Jafarian, M.; Arjomandi, M.; Nathan, G.J. A hybrid solar chemical looping combustion system with a high
solar share. Appl. Energy 2014, 126, 69–77. [CrossRef]
10. Jafarian, M.; Arjomandi, M.; Nathan, G.J. The energetic performance of a novel hybrid solar thermal &
chemical looping combustion plant. Appl. Energy 2014, 132, 74–85.
11. Tanner, J.; Bhattacharya, S. Kinetics of CO2 and steam gasification of Victorian brown coal chars. Chem.
Eng. J. 2016, 285, 331–340. [CrossRef]
12. Patra, T.K.; Sheth, P.N. Biomass gasification models for downdraft gasifier: A state-of-the-art review.
Renew. Sustain. Energy Rev. 2015, 50, 583–593. [CrossRef]
13. Chan, F.L.; Tanksale, A. Review of recent developments in Ni-based catalysts for biomass gasification.
Renew. Sustain. Energy Rev. 2014, 38, 428–438. [CrossRef]
14. Ge, H.; Guo, W.; Shen, L.; Song, T.; Xiao, J. Biomass gasification using chemical looping in a 25kW th reactor
with natural hematite as oxygen carrier. Chem. Eng. J. 2016, 286, 174–183. [CrossRef]
Energies 2019, 12, 2591 13 of 14

15. Adánez, J.; Gayán, P.; Celaya, J.; de Diego, L.F.; García-Labiano, F.; Abad, A. Chemical looping combustion
in a 10 kWth prototype using a CuO/Al2 O3 oxygen carrier: Effect of operating conditions on methane
combustion. Ind. Eng. Chem. Res. 2006, 45, 6075–6080. [CrossRef]
16. Li, F.; Kim, H.R.; Sridhar, D.; Wang, F.; Zeng, L.; Chen, J.; Fan, L.-S. Syngas chemical looping gasification
process: Oxygen carrier particle selection and performance. Energy Fuels 2009, 23, 4182–4189. [CrossRef]
17. Liao, C.; Wu, C.; Yan, Y. The characteristics of inorganic elements in ashes from a 1 MW CFB biomass
gasification power generation plant. Fuel Process. Technol. 2007, 88, 149–156. [CrossRef]
18. Fan, L.; Li, F.; Ramkumar, S. Utilization of chemical looping strategy in coal gasification processes. Particuology
2008, 6, 131–142. [CrossRef]
19. Acharya, B.; Dutta, A.; Basu, P. Chemical-looping gasification of biomass for hydrogen-enriched gas
production with in-process carbon dioxide capture. Energy Fuels 2009, 23, 5077–5083. [CrossRef]
20. Anheden, M.; Svedberg, G. Exergy analysis of chemical-looping combustion systems. Energy Convers. Manag.
1998, 39, 1967–1980. [CrossRef]
21. Zevenhoven-Onderwater, M.; Backman, R.; Skrifvars, B.-J.; Hupa, M. The ash chemistry in fluidised bed
gasification of biomass fuels. Part I: Predicting the chemistry of melting ashes and ash–bed material
interaction. Fuel 2001, 80, 1489–1502. [CrossRef]
22. Florin, N. Calcium looping technologies for gasification and reforming. In Calcium and Chemical Looping
Technology for Power Generation and Carbon Dioxide (CO2 ) Capture; Woodhead Publishing: Sawston, UK, 2015.
23. Matsumura, Y.; Minowa, T.; Potic, B.; Kersten, S.R.A.; Prins, W.; van Swaaij, W.P.M.; van de Beld, B.;
Elliott, D.C.; Neuenschwander, G.G.; Kruse, A. Biomass gasification in near-and super-critical water: Status
and prospects. Biomass Bioenergy 2005, 29, 269–292. [CrossRef]
24. Kruse, A. Supercritical water gasification. Biofuels Bioprod. Biorefining Innov. Sustain. Econ. 2008, 2, 415–437.
[CrossRef]
25. Williams, P.T.; Onwudili, J. Subcritical and supercritical water gasification of cellulose, starch, glucose, and
biomass waste. Energy Fuels 2006, 20, 1259–1265. [CrossRef]
26. Withag, J.A.; Smeets, J.R.; Bramer, E.A.; Brem, G. System model for gasification of biomass model compounds
in supercritical water–a thermodynamic analysis. J. Supercrit. Fluids 2012, 61, 157–166. [CrossRef]
27. Guan, Q.; Savage, P.E.; Wei, C. Gasification of alga Nannochloropsis sp. in supercritical water. J. Supercrit.
Fluids 2012, 61, 139–145. [CrossRef]
28. Yakaboylu, O.; Albrecht, I.; Harinck, J.; Smit, K.; Tsalidis, G.-A.; Di Marcello, M.; Anastasakis, K.; de Jong, W.
Supercritical water gasification of biomass in fluidized bed: First results and experiences obtained from TU
Delft/Gensos semi-pilot scale setup. Biomass Bioenergy 2018, 111, 330–342. [CrossRef]
29. Guo, L.; Lu, Y.; Zhang, X.; Ji, C.; Guan, Y.; Pei, A. Hydrogen production by biomass gasification in supercritical
water: A systematic experimental and analytical study. Catal. Today 2007, 129, 275–286. [CrossRef]
30. Yu, D.; Aihara, M.; Antal, M.J., Jr. Hydrogen production by steam reforming glucose in supercritical water.
Energy Fuels 1993, 7, 574–577. [CrossRef]
31. Yong, T.L.-K.; Matsumura, Y. Reaction kinetics of the lignin conversion in supercritical water. Ind. Eng.
Chem. Res. 2012, 51, 11975–11988. [CrossRef]
32. Zhu, C.; Wang, R.; Jin, H.; Lian, X.; Guo, L.; Huang, J. Supercritical water gasification of glycerol and glucose
in different reactors: The effect of metal wall. Int. J. Hydrog. Energy 2016, 41, 16002–16008. [CrossRef]
33. Reddy, S.N.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of biomass for hydrogen
production. Int. J. Hydrogen Energy 2014, 39, 6912–6926. [CrossRef]
34. Tang, H.; Kitagawa, K. Supercritical water gasification of biomass: Thermodynamic analysis with direct
Gibbs free energy minimization. Chem. Eng. J. 2005, 106, 261–267. [CrossRef]
35. Castello, D.; Fiori, L. Supercritical water gasification of biomass: Thermodynamic constraints. Bioresour.
Technol. 2011, 102, 7574–7582. [CrossRef] [PubMed]
36. Lu, Y.; Zhao, L.; Han, Q.; Wei, L.; Zhang, X.; Guo, L.; Wei, J. Minimum fluidization velocities for supercritical
water fluidized bed within the range of 633–693 K and 23–27 MPa. Int. J. Multiph. Flow 2013, 49, 78–82.
[CrossRef]
37. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. The relative performance of alternative oxygen
carriers for liquid chemical looping combustion and gasification. Int. J. Hydrogen Energy 2017, 42, 16396–16407.
[CrossRef]
Energies 2019, 12, 2591 14 of 14

38. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. Potential of molten lead oxide for liquid chemical
looping gasification (LCLG): A thermochemical analysis. Int. J. Hydrogen Energy 2018, 43, 4195–4210.
[CrossRef]
39. Sarafraz, M.M.; Jafarian, M.; Arjomandi, M.; Nathan, G.J. The thermo-chemical potential liquid chemical
looping gasification with bismuth oxide. Int. J. Hydrogen Energy 2019, 44, 8038–8050. [CrossRef]
40. Salari, E.; Peyghambarzadeh, M.; Sarafraz, M.M.; Hormozi, F. Boiling heat transfer of alumina nano-fluids:
Role of nanoparticle deposition on the boiling heat transfer coefficient. Period. Polytech. Chem. Eng. 2016, 60,
252–258. [CrossRef]
41. Salari, E.; Peyghambarzadeh, S.M.; Sarafraz, M.M.; Hormozi, F.; Nikkhah, V. Thermal behavior of aqueous
iron oxide nano-fluid as a coolant on a flat disc heater under the pool boiling condition. Heat Mass Transf.
2017, 53, 265–275. [CrossRef]
42. Sarafraz, M.M.; Arya, A.; Nikkhah, V.; Hormozi, F. Thermal performance and viscosity of biologically
produced silver/coconut oil nanofluids. Chem. Biochem. Eng. Q. 2017, 30, 489–500. [CrossRef]
43. Arya, A.; Sarafraz, M.M.; Shahmiri, S.; Madani, S.A.H.; Nikkhah, V.; Nakhjavani, S.M. Thermal performance
analysis of a flat heat pipe working with carbon nanotube-water nanofluid for cooling of a high heat flux
heater. Heat Mass Transf. 2018, 54, 985–997. [CrossRef]
44. Sarafraz, M.M.; Arjomandi, M. Demonstration of plausible application of gallium nano-suspension in
microchannel solar thermal receiver: Experimental assessment of thermo-hydraulic performance of
microchannel. Int. Commun. Heat Mass Transf. 2018, 94, 39–46. [CrossRef]
45. Guo, Y.; Wang, S.Z.; Xu, D.H.; Gong, Y.M.; Ma, H.H.; Tang, X.Y. Review of catalytic supercritical water
gasification for hydrogen production from biomass. Renew. Sustain. Energy Rev. 2010, 14, 334–343. [CrossRef]
46. Chakinala, A.G.; Brilman, D.W.F.; van Swaaij, W.P.M.; Kersten, S.R.A. Catalytic and non-catalytic supercritical
water gasification of microalgae and glycerol. Ind. Eng. Chem. Res. 2009, 49, 1113–1122. [CrossRef]
47. Yanik, J.; Ebale, S.; Kruse, A.; Saglam, M.; Yüksel, M. Biomass gasification in supercritical water: II. Effect of
catalyst. Int. J. Hydrogen Energy 2008, 33, 4520–4526. [CrossRef]
48. Aziz, M. Integrated supercritical water gasification and a combined cycle for microalgal utilization. Energy
Convers. Manag. 2015, 91, 140–148. [CrossRef]
49. Wan, W. An innovative system by integrating the gasification unit with the supercritical water unit to produce
clean syngas: Effects of operating parameters. Int. J. Hydrogen Energy 2016, 41, 14573–14582. [CrossRef]
50. Wan, W. An innovative system by integrating the gasification unit with the supercritical water unit to produce
clean syngas for solid oxide fuel cell (SOFC): System performance assessment. Int. J. Hydrogen Energy 2016,
41, 22698–22710. [CrossRef]
51. Li, N.; Li, Y.; Ban, Y.; Song, Y.; Zhi, K.; Teng, Y.; He, R.; Zhou, H.; Liu, Q.; Qi, Y. Direct production of high
hydrogen syngas by steam gasification of Shengli lignite/chars: Remarkable promotion effect of inherent
minerals and pyrolysis temperature. Int. J. Hydrogen Energy 2017, 42, 5865–5872. [CrossRef]
52. Calzavara, Y.; Joussot-Dubien, C.; Boissonnet, G.; Sarrade, S. Evaluation of biomass gasification in supercritical
water process for hydrogen production. Energy Convers. Manag. 2005, 46, 615–631. [CrossRef]
53. McKendry, P. Energy production from biomass (part 1): Overview of biomass. Bioresour. Technol. 2002, 83,
37–46. [CrossRef]
54. Indrawan, N.; Thapa, S.; Bhoi, P.R.; Huhnke, R.L.; Kumar, A. Engine power generation and emission performance
of syngas generated from low-density biomass. Energy Convers. Manag. 2017, 148, 593–603. [CrossRef]
55. Ong, Z.; Cheng, Y.; Maneerung, T.; Yao, Z.; Tong, Y.W.; Wang, C.H.; Dai, Y. Co-gasification of woody biomass
and sewage sludge in a fixed-bed downdraft gasifier. AIChE J. 2015, 61, 2508–2521. [CrossRef]
56. Zhai, Y.; Peng, C.; Xu, B.; Wang, T.; Li, C.; Zeng, G.; Zhu, Y. Hydrothermal carbonisation of sewage sludge for
char production with different waste biomass: Effects of reaction temperature and energy recycling. Energy
2017, 127, 167–174. [CrossRef]
57. Bustamante, F.; Enick, R.; Rothenberger, K.; Howard, B.; Cugini, A.; Ciocco, M.; Morreale, B. Kinetic study of
the reverse water gas shift reaction in high-temperature, high pressure homogeneous systems. Fuel Chem.
Div. Prepr. 2002, 47, 663–664.

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

S-ar putea să vă placă și